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

Dual-gate transistor amplifier in a multimode optomechanical system

Open Access Open Access

Abstract

We present a dual-gate optical transistor based on a multimode optomechanical system, composed of three indirectly coupled cavities and an intermediate mechanical resonator pumped by a frequency-matched field. In this system, two cavities driven on the red mechanical sidebands are regarded as input/ouput gates/poles and the third one on the blue sideband as a basic/control gate/pole, while the resonator as the other basic/control gate/pole. As a nonreciprocal scheme, the significant unidirectional amplification can be resulted by controlling the two control gates/poles. In particular, the nonreciprocal direction of the optical amplification/rectification can be controlled by adjusting the phase differences between two red-sideband driving fields (the pumping and probe fields). Meanwhile, the narrow window that can be analyzed by the effective mechanical damping rate, arises from the extra blue-sideband cavity. Moreover, the tunable slow/fast light effect can be observed, i.e, the group velocity of the unidirectional transmission can be controlled, and thus the switching scheme of slow/fast light effect can also utilized to realize both slow and fast lights through opposite propagation directions, respectively. Such an amplification transistor scheme of controllable amplitude, direction and velocity may imply exciting opportunities for potential applications in photon networks and quantum information processing.

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

1. Introduction

As a valuable phenomenon in various fields of optics, optical non-reciprocity, of which the key element of nonreciprocal devices is the time-reversal symmetry breaking [1,2], provides the possibility that signals can exhibit different transport behaviors along opposite directions and can be used to prevent the unwanted or useless signals getting into the network, thus playing an essential role in the quantum information processing and information protection. Nowadays, nonreciprocal components such as isolators, circulators and directional amplifiers are emerging [36]. The relatively mature ways to implement nonreciprocal devices include magneto-optical effects [710], angular momentum biasing [1113], dynamic modulation [1417], chiral light-matter interaction [18,19], optical nonlinearity [2022]. More surprising, the nonreciprocal effect can also be realized with topological energy transfer by encircling exceptional points, where the adiabatic topological operations is used to transfer energy between mechanical modes [23].

On the other hand, in optomechanical systems, the interaction achieved between optical and mechanical degrees of freedom provides a promising platform for studying quantum sciences at the macroscopic scale [24]. Up to now, a host of breakthroughs have been obtained in optomechanics, such as ground-state cooling of mechanical resonators [25,26], ultralow threshold chaos [27,28], quantum synchronization [2931] and optomechanically induced transparency (OMIT) [3234]. In particular, based on the optomechanical system, the nonreciprocal effect, which can be realized via the quantum interference [3537], Sagnac effect [3841], reservoir engineering [42,43], and so on, has come into view of researchers. The non-reciprocity induced by optomechanics can be modulated by the mechanical driving, the coupling between cavities and so on, of which the relevant schemes have been proposed theoretically [35,4446] and demonstrated experimentally [4750]. Under optomechanical nonreciprocal conditions, directional amplification can be realized with blue-sideband lasers [45,46], coherent mechanical driving [51,52], and active objects (gain) [53,54]. In addition, the optomechanical system is also applied to realize the tunable slow/fast light effect by adjusting relevant parameters [5557] and narrow response windows [58,59].

Motivated by seminal works, we propose a tunable dual-gate optical transistor scheme of directional amplification, which is made up of three indirectly coupled cavities and an intermediate mechanical resonator. In this system, by driving one (two) cavity mode(s) on the blue (red) mechanical sideband and pumping the mechanical resonator with a frequency-matched field, the dual-gate optical transistor will be achieved, where the blue-sideband cavity and the pumped mechanical resonator serve as two gates/poles controlling the photon transmission. We reveal that directional amplification can be achieved and easily controlled by phase modulation. Moreover, the width of the spectral window and the nonreciprocal slow/fast light can be controlled by the blue-sideband driving field. This work provides a new method of controlling the amplification in optomechanical systems, which may inspire follow-up works.

2. Model and transmission coefficient

As shown in Fig. 1(a), we consider an optomechanical system, where three cavity modes $a_k$ ($k= 1, 2, 3$) of frequency $\omega _k$ without inter-coupling, respectively, couple to a common mechanical resonator $b$ of frequency $\omega _m$ via radiation-pressure forces. The red sideband control fields of frequencies $\omega _{c_{1}}$ and $\omega _{c_{2}}$ drive cavities $a_{1}$ and $a_{2}$, respectively, and a weak field of frequency $\omega _{p}$ (amplitude $\varepsilon _{p}$) probes either of the two cavities. As an auxiliary cavity, $a_{3}$ is only driven on the blue sideband at frequency $\omega _{c_{3}}$. A mechanical driving field of frequency $\omega _{b}$ (amplitude $\varepsilon _{b}$) is applied to the resonator $b$ . A kind of implementation scheme is depicted in Fig. 1(b), and such a model can also be realized based on other experimental systems, such as microwave superconducting circuits [48,60] and optomechanical crystal systems [61]. Figure 1(c) shows that there are two closed-loop structures so that relevant transition frequencies should satisfy the frequency-matching condition. The system Hamiltonian has the form $(\hbar =1)$,

$$H=H_{0}+H_{I}+H_{d}.$$

 figure: Fig. 1.

Fig. 1. (a) Schematic diagram of an optomechanical system composed of three indirectly coupled cavities and a mechanical resonator. (b) Schematic illustration of this system composed of three Fabry-Pérot-type resonators with a common movable mirror driven by a mechanical field. (c) Transition relation diagram of internal states in this multi-mode system where $n_{ak}$ and $n_{b}$ are the excitation numbers of the three cavity modes and the mechanical mode, respectively.

Download Full Size | PDF

Here $H_{0}$ represents the free Hamiltonian of the system,

$$H_{0}=\sum_{k=1}^{3}\omega_{k}a_{k}^{\dagger}a_{k}+\omega_{m}b^{\dagger}b;$$
$H_{I}$ describes the interaction between the cavity $a_k$ and the mechanical mode $b$ with the single-photon optomechanical coupling strength $g_{k}$,
$$\begin{aligned} H_{I} = &-g_{1}a_{1}^{\dagger}a_{1}(b^{\dagger}+b)+g_{2}a_{2}^{ \dagger}a_{2}(b^{\dagger}+b)\\ &\,+g_{3}a_{3}^{\dagger}a_{3}(b^{\dagger}+b); \end{aligned}$$
$H_{d}$ denotes the driving term including control and probe fields on the cavities and the resonator,
$$\begin{align*} H_{d} = &\sum_{k=1}^{3}\varepsilon_{ck}(ia_{k}^{\dagger}e^{-i \omega_{ck}t}e^{i\theta_{k}}+H.c.)\\ &+\varepsilon_{p}(ia_{1}^{\dagger}e^{-i\omega_{p}t}+H.c.)\\ &+\varepsilon_{b}(ib^{\dagger}e^{-i\omega_{b}t}+H.c.) , \end{align*}$$
with $\varepsilon _{ck}$ ($\theta _{k}$) representing the real amplitude (phase) of the control field $\varepsilon _{ck}$.

In the rotating frame with respect to the control field frequency, respectively, the Heisenberg-Langevin equations for the operators in the system can be obtained as [62],

$$\begin{aligned} \dot{a_{1}} = &-(\kappa+i\Delta_{1})a_{1}+ig_{1}a_{1}(b+b^{\dagger})+ \varepsilon_{c1}e^{i\theta_{1}}\\ &+\varepsilon_{p}e^{-i\sigma t}+\sqrt{2\kappa}a_{1}^{in},\\ \dot{a_{2}} = &-(\kappa+i\Delta_{2})a_{2}-ig_{2}a_{2}(b+b^{\dagger})+ \varepsilon_{c2}e^{i\theta_{2}}\\ &+\sqrt{2\kappa}a_{2}^{in},\\ \dot{a_{3}} = &-(\kappa+i\Delta_{3})a_{3}-ig_{3}a_{3}(b+b^{\dagger})+ \varepsilon_{c3}e^{i\theta_{3}}\\ &+\sqrt{2\kappa}a_{3}^{in},\\ \dot{b} = &-(\gamma_{m}+i\omega_{m})b+i(g_{1}a_{1}^{\dagger}a_{1}-g_{2}a_{2}^{ \dagger}a_{2}-g_{3}a_{3}^{\dagger}a_{3})\\ &+\varepsilon_{b}e^{-i\omega_{b}t}+\sqrt{2\gamma_{m}}b^{in}, \end{aligned}$$
where $\sigma =\omega _{p}-\omega _{ck}$ ($\Delta _{k}=\omega _{k}-\omega _{ck}$) is the detuning of the probe field (cavity) from the control field and $\sigma =\omega _{b}$ for frequency match.

Taking relevant dissipations and noises in account, we introduce the damping rate $\kappa$ ($\gamma _{m}$) of the intrinsic cavity (resonator) and the noise operator $a_{k}^{in}$ ($b^{in}$) of zero mean value associated with $\kappa$ ($\gamma _{m}$). Aiming to linearization, each operator is written as the sum of its steady-state value and a small fluctuation, i.e., $a_{k}=\langle a_{k}\rangle +\delta a_{k}$ and $b=\langle b\rangle +\delta b$. Afterwards, the steady-state solutions are written as,

$$\begin{aligned} \langle a_{1}\rangle &= \frac{\varepsilon_{c1}e^{i\theta_{1}}}{\kappa+i\Delta_{1}^{ \prime }},\\ \langle a_{2}\rangle &= \frac{\varepsilon_{c2}e^{i\theta_{2}}}{\kappa+i\Delta_{2}^{ \prime }},\\ \langle a_{3}\rangle &= \frac{\varepsilon_{c3}e^{i\theta_{3}}}{\kappa+i\Delta_{3}^{ \prime }},\\ \langle b\rangle &= \frac{i(g_{1}|\alpha_{1}|^{2}-g_{2}|\alpha_{2}|^{2}-g_{3}| \alpha_{3}|^{2})}{\gamma_{m}+i\omega_{m}}, \end{aligned}$$
with $\Delta _{1}^{\prime }=\Delta _{1}-g_{1}[\langle b\rangle +\langle b\rangle ^{*}]$, $\Delta _{2,3}^{\prime }=\Delta _{2,3}+g_{2,3}[\langle b\rangle +\langle b\rangle ^{*}]$.

By neglecting the nonlinear terms in Eq. (5), a series of linearized equations can be obtained,

$$\begin{aligned} \delta\dot{a_{1}} = &-(\kappa+i\Delta_{1}^{\prime })\delta a_{1}+iG_{1}(\delta b+\delta b^{\dagger})+\varepsilon_{p}e^{-i\sigma t}\\ &+\sqrt{2\kappa}a_{1}^{in},\\ \delta\dot{a_{2}} = &-(\kappa+i\Delta_{2}^{\prime })\delta a_{2}-iG_{2}(\delta b+\delta b^{\dagger})+\sqrt{2\kappa}a_{2}^{in},\\ \delta\dot{a_{3}} = &-(\kappa+i\Delta_{3}^{\prime })\delta a_{3}-iG_{3}(\delta b+\delta b^{\dagger})+\sqrt{2\kappa}a_{3}^{in},\\ \delta\dot{b} = &-(\gamma_{m}+i\omega_{m})\delta b+i(G_{1}^{*}\delta a_{1}+G_{1}\delta a_{1}^{\dagger}-G_{2}^{*}\delta a_{2}\\ &-G_{2}\delta a_{2}^{\dagger}-G_{3}^{*}\delta a_{3}-G_{3}\delta a_{3}^{\dagger})+\varepsilon_{b}e^{-i\omega_{b}t}\\ &+\sqrt{2\gamma_{m}}b^{in}, \end{aligned}$$
where $G_{k}=g_{k}\langle a_{k}\rangle$ represents the effective linear coupling strength between the cavity and the mechanical mode. Here $\omega _{m} = \Delta _{1}^{\prime } = \Delta _{2}^{\prime } = -\Delta _{3}^{\prime }$ owing to cavities $a_{1}$ and $a_{2}$ driven on the red sidebands and $a_{3}$ on the blue one.

Assuming that the system is operated in the resolved sideband regime with $\omega _{m}\gg \kappa , \gamma _{m}, G_{k}$, we do the following substitution: $\delta a_{k}\rightarrow \delta a_{k}e^{-i\Delta _{k}^{\prime }t}$, $\delta a_{k}^{in}\rightarrow \delta a_{k}^{in}e^{-i\Delta _{k}^{\prime }t}$, $\delta b\rightarrow \delta be^{-i\omega _{m}t}$, $\delta b^{in}\rightarrow \delta b^{in}e^{-i\omega _{m}t}$ [35,51]. Equation (7) becomes,

$$\begin{aligned} \delta\dot{a_{1}} = &-\kappa\delta a_{1}+iG_{1}\delta b+\varepsilon_{p}e^{-i\delta t}+\sqrt{2\kappa}a_{1}^{in},\\ \delta\dot{a_{2}} = &-\kappa\delta a_{2}-iG_{2}\delta b+\sqrt{2\kappa} a_{2}^{in},\\ \delta\dot{a_{3}}^{\dagger} = &-\kappa\delta a_{3}^{\dagger}+iG_{3}^{*}\delta b+\sqrt{2\kappa}a_{3}^{in\dagger},\\ \delta\dot{b} = &-\gamma_{m}\delta b+iG_{1}^{*}\delta a_{1}-iG_{2}^{*}\delta a_{2}-iG_{3}\delta a_{3}^{\dagger}\\ &+\varepsilon_{b}e^{-i\delta t}+\sqrt{2\gamma_{m}}b^{in}, \end{aligned}$$
where $\delta =\sigma -\Delta _{1}^{\prime }=\omega _{b}-\omega _{m}$.

To solve Eq. (8), the equations are transformed into another interaction picture with $\delta a_{k}\rightarrow \delta a_{k}e^{-i\delta t}, \delta b\rightarrow \delta be^{-i\delta t}$. Neglecting the noise terms, substituting gives

$$\begin{aligned} \delta\dot{a_{1}} = &-(\kappa-i\delta)\delta a_{1}+iG_{1}\delta b+\varepsilon_{p},\\ \delta\dot{a_{2}} = &-(\kappa-i\delta)\delta a_{2}-iG_{2}\delta b,\\ \delta\dot{a_{3}}^{\dagger} = &-(\kappa-i\delta)\delta a_{3}^{\dagger}+iG_{3}^{*}\delta b,\\ \delta\dot{b} = &-(\gamma_{m}-i\delta)\delta b+iG_{1}^{*}\delta a_{1}-iG_{2}^{*}\delta a_{2}\\ &-iG_{3}\delta a_{3}^{\dagger}+\varepsilon_{b}. \end{aligned}$$
For simplicity, we assume $G_{1}=G$ to be real, $G_{2}=Ge^{-i\phi }$ with a phase difference $\phi$ from $G_{1}$, and then, by solving the steady state of Eq. (9), we obtain
$$\begin{aligned} \langle\delta a_{1}\rangle &= \frac{-|G|^{2}\varepsilon_{p}+iG\varepsilon_{b}( \kappa-i\delta)}{N}+\frac{\varepsilon_{p}}{\kappa-i\delta},\\ \langle\delta a_{2}\rangle &= \frac{G^{2}e^{-i\phi} \varepsilon_{p}-iGe^{-i\phi}\varepsilon_{b}(\kappa-i\delta)}{N} \end{aligned}$$
with the denominator $N=(\gamma -i\delta )(\kappa -i\delta )^{2}+(\kappa -i \delta )(2|G|^{2}-|G_{3}|^{2})$.

Then the cavity output fields $\langle \delta a_{1,2}^{out}\rangle$ can be obtained by the standard input-output relation $\langle \delta a_{1,2}^{out}\rangle +\langle \delta a_{1,2}^{in}\rangle =\sqrt { 2\kappa }\langle \delta a_{1,2}\rangle$ [63] in the overcoupling regime [64,65]. When $\langle \delta a_{1}^{in}\rangle =\varepsilon _{p}/\sqrt {2\kappa }$ and $\langle \delta a_{2}^{in}\rangle =0$, the transmission coefficient $t_{21}\equiv \partial \langle \delta a_{2}^{out}\rangle /\partial \langle \delta a_{1}^{in}\rangle$ [51,66] can be derived

$$t_{21}=\frac{2\kappa[G^{2}e^{-i\phi}-iGye^{-i(\phi+\beta)}(\kappa-i\delta)]}{ N}$$
with $y=|\varepsilon _{b}/\varepsilon _{p}|$ and the phase difference $\beta$ between $\varepsilon _{b}$ and $\varepsilon _{p}$. In a similar manner, when only a weak field probes from $a_{2}$ to $a_{1}$, the transmission coefficient is
$$t_{12}=\frac{2\kappa[G^{2}e^{i\phi}-iGye^{-i\beta}(\kappa-i\delta)]}{N}.$$
Then, the nonreciprocal optical transmission can be observed with the nonzero phase difference $\phi$ by the corresponding transmissivity of $T_{21}\equiv |t_{21}|^{2}$ ($a_{1}\rightarrow a_{2}$) and $T_{12}\equiv |t_{12}|^{2}$ ($a_{2}\rightarrow a_{1}$). Breaking the time-reversal symmetry, we may obtain undirectional propagation of the probe fields. Noteworthily, it is needed that a frequency-matched drive field to drive the mechanical resonator, and the cavity driven on blue sideband is like a bath of the gain.

On account of the gain arising from the blue-sideband laser, the stability of the system should be tested. The coupled Eq. (8) can be solved self-consistently, and then they can be rewritten into the matrix form

$$\dot{\tau}=M\tau+\nu+\mu,$$
with the vector $\tau =(\delta a_{1},\delta a_{2},\delta a_{3}^{\dagger },\delta b)^{T}$, $\nu =(\varepsilon _{p}e^{-i\delta t}, 0, 0, \varepsilon _{b}e^{-i\delta t})^{T}$, $\mu =(\sqrt {2\kappa }a_{1}^{in},\sqrt { 2\kappa }a_{2}^{in},\sqrt {2\kappa } a_{3}^{in\dagger },\sqrt {2\gamma _{m}} b^{in})^{T}$, and the coefficient matrix
$$M= \begin{pmatrix} -\kappa & 0 & 0 & iG_{1} \\ 0 & -\kappa & 0 & -iG_{2} \\ 0 & 0 & -\kappa & iG_{3}^{*} \\ iG_{1}^{*} & -iG_{2}^{*} & -iG_{3} & -\gamma_{m} \ \end{pmatrix}.$$
The Routh-Hurwitz criterion can be applied to derive the stability condition [67], whereas it is too cumbersome to give here [46]. Note that, the situation that eigenvalues of matrix $M$ have no positive real parts indicates the stability of the system. With the expression of $\langle a_{k}\rangle$ in Eq. (6), the coupling strength $G_{k}$ can be adjusted via phase modulation of phase $\theta _{k}$. Consider $G_{3}=nG(n\geq 0)$ with a adjustable ratio $n$, for investigating the parameter range for the stability. Figure 2 shows the stability diagrams of the system. According to Fig. 2(a) with $\phi$, obviously, $n\;<\;1.413$ (n=1.413 denotes the boundary between stable and unstable regions) should be guaranteed for the stability. It is found that there is no change of the stable regions with changing $\phi$, i.e., the stability is independent of $\phi$. That can be demonstrated further by Fig. 2(b), showing that the stable region covers the whole $\phi$. Relevant parameters for stability will be used in the following discussion.

 figure: Fig. 2.

Fig. 2. (a) Stability diagram of the system on the plane of coupling strengths $G_{1}$ and $G_{3}$ for $\phi =0$ (The diagram is the same as those with $\phi$ of any value.). (b) Stability diagram of the system on the plane of phase difference $\phi$ and the coupling strength $G_{1}$ with $n=2$. The other parameter is $\gamma _{m}=\kappa /100$. The yellow regions are stable and the green regions are unstable.

Download Full Size | PDF

3. Results and discussion

In this section, we will study optical responses of the system as a dual-gate optical transistor with amplification due to the gain arising from the auxiliary cavity $a_{3}$ on blue sideband and the mechanical resonator $b$ pumped by field $\varepsilon _{b}$ [51], including the manipulation of the direction (nonreciprocal amplification and unidirectionality) and group velocity (slow/fast light effect) of photon transport. There being no need for the frequency matching between cavities due to their decoupling, the cavity can be either microwave or optical one. Thus, the modulation of two photon flows, two microwave fields or both of them all can be realized. Clearly, such a scheme proposed is an open platform. Here we regard the cavity $a_{1}$ ($a_{2}$ and $a_{3}$) in the microwave (optical) domain. Thus, our system shows the ability to convert the frequencies of input signals under the frequency-mixing process, e.g., when inputting an optical signal into $a_{1}$, the retrieval signal from $a_{2}$ can be microwave with proper parameters. Similar to electricity, the cavity can be treated as a gate/grid of the transistor (like electrical triode), where the base poles are $a_{3}$ and $b$ for controlling the amplification and other gates are $a_{1}$ and $a_{2}$.

The undirectional amplification, which can be investigated by comparing two transmissivities $T_{12}$ and $T_{21}$, may be more obvious when $a_{3}$ is driven by a blue-sideband field, and can be modulated by adjusting the coupling strength $G_{3}$ from Eqs. (11) and (12). Figure 3(a) shows that $T_{12}$ is very sensitive to $G_{3}$, i.e., it is enhanced obviously with the increasing $G_{3}$ around resonant point $\delta =0$. However, Fig. 3(b) shows that $T_{21}$ without obvious gain around resonance becomes zero at $\delta =0$. That means the transmission of $a_{1}\rightarrow a_{2}$ is suppressed completely at the resonance point. Thus, the nonreciprocity between $T_{12}$ and $T_{21}$ can be induced in the near-resonance range. For the sake of clarity, in Fig. 3(c), we plot $T_{12}$ and $T_{21}$ as a function of $\delta$ with $G_{3}=2.8\kappa$ (in the stable region). It is found that, at the resonance point, the transmission of $a_{1}\rightarrow a_{2}$ is suppressed but the opposite transmission is amplified near $10^{4}$ times. Obviously, there is a strong unidirectional amplification for the probe field in the vicinity of $\delta =0$. Note the blue-sideband field driving $a_{3}$ may play a important role in this unidirectional amplification. For comparison, in Fig. 3(d) the nonreciprocity of the system with $a_{3}$ on the red sideband will also be observed, whose tranmissivities can be given just with the substitution $-|G_{3}|^{2}\rightarrow +|G_{3}|^{2}$ in the denominator $N$. With the same $G_{3}$, there is no gain for both transmission at the resonance point. The physics mechanism of the amplification is that the cavity being driven by the blue-sideband field will induce a anti-Stokes process in this system, which can be viewed as a heat bath. The unidirectional amplification is tunable by modulating $G_{3}$, thus the cavity $a_{3}$ can be regarded as one basic pole/gate of the optical transistor for controlling the amplification magnitude.

 figure: Fig. 3.

Fig. 3. Logarithms of transmissivities (a) $T_{12}$ and (b) $T_{21}$ versus detuning $\delta$ and coupling strength $G_{3}$ for cavity $a_{3}$ on blue sideband with $y=2$. Logarithms of the transmissivities $T_{12}$ and $T_{21}$ with $G_{3}=2.8\kappa$ and $y=2$ (c) for $a_{3}$ on blue sideband; (d) for $a_{3}$ on red sideband. The transmissivity $T_{12}$ versus $\delta$ with (e) $G_{3}=0$ and (f) $G_3=2.8\kappa$ for cavity $a_{3}$ on blue sideband. Other parameters are $\phi =\pi$, $\beta =\pi /2$, $\gamma _{m}=\kappa /100$ and $G=2\kappa$.

Download Full Size | PDF

In addition, another effect of the basic pole/gate $a_{3}$ is narrowing the linewidth of the spectral window by increasing $G_{3}$, still caused by its operation in the blue-sideband domain. As shown in Figs. 3(e) and 3(f), increasing $G_{3}$ from 0 to $2.8\kappa$ with $a_{3}$ on the blue sideband, the linewidth of the window becomes narrow, where $G_3=0$ corresponds to the common three-mode optomechanical system. It is clear that the number of peaks is tunable by the coupling strength. In our system, the increasing of $G_3$ merges the three peaks into only one with the narrower window. The phenomenon can be well explained by the effective mechanical damping rate determining the linewidth of the transparency (amplification or absorption) window in optomechanics [24,58,59]. Through the steady-state solution of $\langle \delta b\rangle$ in Eq. (9),

$$(\gamma_m-i\delta+\frac{|G_{1}|^{2}+|G_{2}|^{2}-|G_{3}|^{2}}{\kappa-i\delta})\langle\delta b\rangle=(\frac{iG_{1}^{*}}{\kappa-i\delta}+ye^{-i\beta})\varepsilon_{p},$$
in which the real part of the coefficient of $\langle \delta b\rangle$ represents the effective mechanical damping rate and the imaginary part represents the effective frequency, i.e., the damping rate $\gamma _{m}$ and detuning $\delta$ are modulated by a complex quantity $\frac {|G_{1}|^{2}+|G_{2}|^{2}-|G_{3}|^{2}}{\kappa -i\delta }$ , respectively. Then, we can obtain the effective mechanical damping rate with $y\;<<\;1$ as
$$\Gamma_{b}=\gamma_{m}+\frac{\kappa(2|G|^{2}-|G_{3}|^{2})}{\kappa^{2}+\delta^{2}},$$
which is appropriate for $a_{3}$ on the blue sideband. As for red-sideband case, just turn $-|G_{3}|^{2}$ to $+|G_{3}|^{2}$,
$$\Gamma_{r}=\gamma_{m}+\frac{\kappa(2|G|^{2}+|G_{3}|^{2})}{\kappa^{2}+\delta^{2}}.$$
Comparing Eq. (16) and Eq. (17), obviously, $\Gamma _{r}\;>\;\Gamma _{G_{3}=0}\;>\;\Gamma _{b}$ is surely satisfied, i.e., the linewidth of the blue-sideband case is much smaller than that of the red-sideband case or the case without the third cavity. The window is getting narrower with the increaseing of $G_3$ according to Eq. (16). Therefore, the window linewidth can be controlled by the auxiliary cavity $a_{3}$, and the narrow windows may have important applications, such as precise measurements and operations [68,69].

As the other basic pole/gate of this optical transistor amplifier, the driven mechanical resonator can not only control amplification but the direction of the nonreciprocity by the modulation of the phase difference $\beta$, to adjust which we vary the ratio $y$. Figure 4(a) shows logarithms of $T_{21}$ and $T_{12}$ as a function of $y$ with $G_{3}=2.8\kappa$. When $y=0$ corresponding to $\varepsilon _{b}=0$, $T_{21}=T_{12}$ implies the reciprocity. With the increasing of $y$, $T_{12}$ is amplified gradually but $T_{21}$ becomes zero at $y=2$, which shows strong unidirectional amplification. Obviously, the amplification surely gets larger with increasing $y$. With Fig. 1(c), the underlying physics can be expressed. As depicted in Fig. 1(c), such a system of five states involving excitation numbers $n_{ak}$ and $n_{b}$ forms the closed-loop transition structure sensitive to the phase. The transition $|n_{a1}, n_{a2}, n_{a3}+1, n_{b}\rangle \leftrightarrow |n_{a1}, n_{a2}, n_{a3}, n_{b}+1\rangle$ can be regarded to be the association of this structure, thus $a_{3}$ is a basic pole only for controlling the amplification magnitude. In the absence of the transition $|n_{a1}, n_{a2}, n_{a3}, n_{b}\rangle \leftrightarrow |n_{a1}, n_{a2}, n_{a3}, n_{b}+1\rangle$, there is only one loop $|n_{a1}, n_{a2}, n_{a3}, n_{b}\rangle \leftrightarrow |n_{a1}+1, n_{a2}, n_{a3}, n_{b}\rangle \leftrightarrow |n_{a1}, n_{a2}, n_{a3}, n_{b}+1\rangle \leftrightarrow |n_{a1}, n_{a2}+1, n_{a3}, n_{b}\rangle \leftrightarrow |n_{a1}, n_{a2}, n_{a3}, n_{b}\rangle$. When inputting probe field $\varepsilon _{p}$ from $a_{1}$, a phase-matched filed out of $a_{2}$ can arise from the four-wave mixing independent of phase differences, and vice verse. Therefore, the transmissions of $a_{1}\rightarrow a_{2}$ and $a_{2}\rightarrow a_{1}$ are reciprocal. However, in the presence of that transition, two different closed loops made up will show different phase-dependent behaviors, respectively, thus the phase difference introduced plays an essential role in the quantum interference between two transition paths of the probe fields and then optical nonreciprocity. The importance of $\phi$ for the nonreciprocity can also seen from Eqs. (11) and (12).

 figure: Fig. 4.

Fig. 4. (a) Logarithms of the transmissivities $T_{21}$ and $T_{12}$ and (b) isolation ratio $I$ versus relative ratio $y$ with $\delta =0$. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

However, to examine the quality of nonreciprocity (like for a optical switching device), the isolation ratio

$$I=\frac{T_{12}-T_{21}}{T_{12}+T_{21}}$$
is very useful. As shown in Fig. 4(b), $y=0$ and $y=2$ correspond to the minimum ($I=0$) and maximum ($I=1$) optical reciprocity of the devices, respectively. That is to say, the tunable nonreciprocal amplification can be achieved with a certain frequency-matched driving field and the blue-sideband field. So it is reasonable to consider the driven mechanical resonator as another gate/pole off thus transistor amplifier. With the above analysis, the multimode optomechanical system can be in analogy with a dual-gate optical transistor. Compared with Ref. [51], the main advantage is that the auxiliary cavity $a_{3}$ supports a method to control and optimize the unidirectional amplification by adjusting $G_{3}$. To maintain and optimize the unidirectional transmission, $y$ is fixed as discussed in Fig. 4 where $y=2$. Thus, the adjusting of $G_{3}$ may become the only method to control amplification by adjusting the field intensity, which shows the necessity of the existence of $a_{3}$. In the amplifier, $a_{3}$ as the $1st$ gate/pole adjust amplification, and $b$ as the $2nd$ gate/pole mainly tune nonreciprocity. Both contributing to the transistor amplifier, the two gates supply with each other and neither can be omitted.

As mentioned above, the phase is vital and both phase differences $\phi$ and $\beta$ can modulate the transport behavior from Eqs. (11) and (12). In addition, from those equations, the nonreciprocity should need $\phi \neq \pm 2k\pi$ ($k=0,1,2,3\cdots$) to satisfy the Lorentz reciprocal theorem no matter what $\beta$ value. As in Figs. 5(a)–5(b) with $\phi =\pi /2$, the unidirectional amplification can be obtained with the broken time-reversal symmetry of the system. With $\beta =0$ in Fig. 5(a), $log_{10}(T_{21})\simeq 4$ and $T_{12}\simeq 0$ at $\delta =0$ means unidirectional transmission $a_{1}\rightarrow a_{2}$; tuning $\beta$ to $\pi /2$, the direction of unidirectional transmission is reversed. Figures 5(c)–5(d) show that the nonreciprocal transmission behavior is changed by varying $\beta$. With a few certain parameters (like $\beta =0$ and $\phi =\pi$ in Fig. 5(d)), $T_{21}=T_{12}$ appearing indicates the reciprocal transmission. It is noteworthy that, by modulating the phase $\beta$, the direction of the amplified (suppressed) output probe field can be switch, i.e., the direction of amplification of the optical transistor, for expressing which we plot isolation ratio $I$ against $\beta$ in Fig. 5(g). It is more intuitive that the isolation ratio varies with $\beta$ periodically and continuously between 1 and −1, which indicates the two opposite unidirectional amplification, respectively. As shown in Figs. 5(e)–5(f), with a given $\beta$, $T_{21}$ is independent on $\phi$, which is also be deduced from Eq. (11). Thus, $\phi$ is just essential for nonreciprocity but cannot change the direction. Strikingly, when $\beta =\pi /2, \phi =0, \pm 2k\pi$ both probe fields can be suppressed simultaneously, i.e., bidirectional opaque, as in Fig. 5(h). In this regard, we can realize tunable switching scheme among the bidirectional opaque and transmission, and unidirectional transmission with controllable direction.

 figure: Fig. 5.

Fig. 5. Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\delta$ for $\phi =\pi /2$ and (a) $\beta =0$; (b) $\beta =\pi /2$. Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\beta$ for $\delta =0$ and (c) $\phi =\pi /2$; (d) $\phi =\pi$. Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\phi$ for $\delta =0$ and (e) $\beta =3\pi /2$; (f) $\beta =\pi$. (g) Isolation ratio $I$ verse $\beta$ for $\phi =\pi /2$ and $\delta =0$. (h) Transmissivities $T_{21}$ and $T_{12}$ verse $\phi$ for $\beta =\pi /2$. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

Moreover, besides the direction, the group velocity of the probe field in this transistor may be also controlled, i.e., the slow/fast light effect. Here we focus on the tunable fast and slow light effect with undirectional amplification, and the way to the switching between slow and fast lights by adjusting $G_{3}$ arising from the auxiliary cavity $a_{3}$. The optical transmission group delay is defined as [53,56]

$$\tau_{12(21)}=\frac{d\Theta_{12(21)}}{d\omega_{p}},$$
where $\Theta _{12(21)}=arg[t_{12(21)}(\omega _{p})]$ is the phase of the relevant output field at frequency $\omega _{p}$.

In the situation of directional transmission of $T_{12} \neq T_{21}=0$ for $\phi =\beta =\pi /2$, Fig. 6(a) shows the group delay $\tau _{12}$ as a function of the coupling strength $G_{3}$ at the resonance, where $\tau _{12}=0$ with $G_{3}=2.65\kappa$. When $G_{3}\;<\;2.65\kappa$ ($>\;2.65\kappa$), $\tau _{12}\;<\;0$ ($>\;0$) corresponding to the fast (slow) light. That is to say, the light group velocity of the directional transmission can be modulated and switched between slow and fast light effects by adjusting $G_{3}$. Thus, both the amplitude and direction of the directional amplification can be easily controlled, involving the slow/fast light effect in this system. In the situation of bidirectional amplification for $\phi =\pi$ and $\beta =0$, the system may show asymmetric light propagations of different group velocities. Then, near the resonance (in the vicinity of $\delta =-1.1\kappa$), the group velocity $a_{1}\rightarrow a_{2}$ is slowed down, whereas the transmission $a_{2}\rightarrow a_{1}$ shows the fast light. At the symmetric position around $\delta =1.1\kappa$, the transmissions along two directions have inverse fast/slow light behaviors. Generally, slow light can be observed along with the standard OMIT, owing to the anomalous dispersion. Essentially, the fast and slow light effects are physically governed by the anti-rotating-wave and rotating-wave interaction terms [56,57], respectively. There are both anti-rotating-wave (two-mode-squeezing interaction induced by the blue-sideband cavity $a_3$) and rotating-wave (beam-splitter interactions induced by the red-sideband cavities $a_1$ and $a_2$) terms, which may have competition with each other. Therefore, in such a system, we can observe fast/slow light effect and the switch between them. The group velocity difference between two transmissions may imply opportunities for precise measurement.

 figure: Fig. 6.

Fig. 6. (a) Group delay $\tau _{12}$ versus coupling strength $G_{3}$ for $\phi =\pi /2$, $\beta =\pi /2$ and $\delta =0$. (b) Group delays $\tau _{12}$ and $\tau _{21}$ versus detuning $\delta$ for $\phi =\pi$ and $\beta =0$. $\tau _{12(21)}=0$ is the transition point of slow and fast lights. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

4. Conclusions

In summary, we proposed a dual-gate optical transistor in a multimode optomechanical system, composed of three microwave/optical cavities without direct interaction and an intermediate driven mechanical resonator $b$, where cavities $a_{1}$ and $a_{2}$ are driven on red sidebands and cavity $a_{3}$ on blue sideband. Thus, we fabricate a closed-loop mode structure sensitive to the phase. We can adjust the amplitude of the unidirectional amplification by modifying the amplitude of driving fields on $a_{3}$ and $b$, which act as two basic poles/gates of the optical transistor, while the direction of the unidirectional transmission can be turned by phase modulation. The narrower windows induced by the blue-sideband field can be observed in our system. In addition, this transistor has the ability to switch transmission between slow and fast lights and then shows nonreciprocal group velocities. Such an optical transistor amplifier with many prominent functions may provide potential applications for quantum optical devices and has extraordinary promising prospects.

Funding

National Natural Science Foundation of China (11704063, 11704064); Jilin Scientific and Technological Development Program (20180520205JH); Fundamental Research Funds for the Central Universities (2412019FZ045); Science Foundation of the Education Department of Jilin Province during the 13th Five-Year Plan Period (JJKH20180010KJ).

Acknowledgments

Thanks go to Y. Li for fruitful discussions.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. R. J. Potton, “Reciprocity in optics,” Rep. Prog. Phys. 67(5), 717–754 (2004). [CrossRef]  

2. Y. Tokura and N. Nagaosa, “Nonreciprocal responses from non-centrosymmetric quantum materials,” Nat. Commun. 9(1), 3740 (2018). [CrossRef]  

3. K. M. Sliwa, M. Hatridge, A. Narla, S. Shankar, L. Frunzio, R. J. Schoelkopf, and M. H. Devoret, “Reconfigurable josephson circulator/directional amplifier,” Phys. Rev. X 5(4), 041020 (2015). [CrossRef]  

4. I. Liberal and N. Engheta, “Multiqubit subradiant states in n-port waveguide devices: $\epsilon$-and-μ-near-zero hubs and nonreciprocal circulators,” Phys. Rev. A 97(2), 022309 (2018). [CrossRef]  

5. Z. Shen, Y.-L. Zhang, Y. Chen, F.-W. Sun, X.-B. Zou, G.-C. Guo, C.-L. Zou, and C.-H. Dong, “Reconfigurable optomechanical circulator and directional amplifier,” Nat. Commun. 9(1), 1797 (2018). [CrossRef]  

6. B. Abdo, N. T. Bronn, O. Jinka, S. Olivadese, A. D. Córcoles, V. P. Adiga, M. Brink, R. E. Lake, X. Wu, D. P. Pappas, and J. M. Chow, “Active protection of a superconducting qubit with an interferometric josephson isolator,” Nat. Commun. 10(1), 3154 (2019). [CrossRef]  

7. L. Aplet and J. Carson, “A faraday effect optical isolator,” Appl. Opt. 3(4), 544–545 (1964). [CrossRef]  

8. Z. Wang, Y. Chong, J. D. Joannopoulos, and M. Soljačić, “Observation of unidirectional backscattering-immune topological electromagnetic states,” Nature 461(7265), 772–775 (2009). [CrossRef]  

9. Y. Hadad and B. Z. Steinberg, “Magnetized spiral chains of plasmonic ellipsoids for one-way optical waveguides,” Phys. Rev. Lett. 105(23), 233904 (2010). [CrossRef]  

10. A. B. Khanikaev, S. H. Mousavi, G. Shvets, and Y. S. Kivshar, “One-way extraordinary optical transmission and nonreciprocal spoof plasmons,” Phys. Rev. Lett. 105(12), 126804 (2010). [CrossRef]  

11. R. Fleury, D. L. Sounas, C. F. Sieck, M. R. Haberman, and A. Alù, “Sound isolation and giant linear nonreciprocity in a compact acoustic circulator,” Science 343(6170), 516–519 (2014). [CrossRef]  

12. D.-W. Wang, H.-T. Zhou, M.-J. Guo, J.-X. Zhang, J. Evers, and S.-Y. Zhu, “Optical diode made from a moving photonic crystal,” Phys. Rev. Lett. 110(9), 093901 (2013). [CrossRef]  

13. N. A. Estep, D. L. Sounas, J. Soric, and A. Alù, “Magnetic-free non-reciprocity and isolation based on parametrically modulated coupled-resonator loops,” Nat. Phys. 10(12), 923–927 (2014). [CrossRef]  

14. Z. Yu and S. Fan, “Complete optical isolation created by indirect interband photonic transitions,” Nat. Photonics 3(2), 91–94 (2009). [CrossRef]  

15. L. D. Tzuang, K. Fang, P. Nussenzveig, S. Fan, and M. Lipson, “Non-reciprocal phase shift induced by an effective magnetic flux for light,” Nat. Photonics 8(9), 701–705 (2014). [CrossRef]  

16. 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]  

17. S. Horsley, J.-H. Wu, M. Artoni, and G. La Rocca, “Optical nonreciprocity of cold atom bragg mirrors in motion,” Phys. Rev. Lett. 110(22), 223602 (2013). [CrossRef]  

18. S. Yoo and Q.-H. Park, “Chiral light-matter interaction in optical resonators,” Phys. Rev. Lett. 114(20), 203003 (2015). [CrossRef]  

19. S. Mahmoodian, P. Lodahl, and A. S. Sørensen, “Quantum networks with chiral-light–matter interaction in waveguides,” Phys. Rev. Lett. 117(24), 240501 (2016). [CrossRef]  

20. 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]  

21. X. Guo, C.-L. Zou, H. Jung, and H. X. Tang, “On-chip strong coupling and efficient frequency conversion between telecom and visible optical modes,” Phys. Rev. Lett. 117(12), 123902 (2016). [CrossRef]  

22. A. R. Hamann, C. Müller, M. Jerger, M. Zanner, J. Combes, M. Pletyukhov, M. Weides, T. M. Stace, and A. Fedorov, “Nonreciprocity realized with quantum nonlinearity,” Phys. Rev. Lett. 121(12), 123601 (2018). [CrossRef]  

23. H. Xu, D. Mason, L. Jiang, and J. G. E. Harris, “Topological dynamics in an optomechanical system with highly non-degenerate modes,” arXiv:1703.07374 (2017).

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

25. J. Chan, T. 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]  

26. P. Stadler, W. Belzig, and G. Rastelli, “Ground-state cooling of a carbon nanomechanical resonator by spin-polarized current,” Phys. Rev. Lett. 113(4), 047201 (2014). [CrossRef]  

27. 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]  

28. G.-L. Zhu, X.-Y. Lü, L.-L. Zheng, Z.-M. Zhan, F. Nori, and Y. Wu, “Single-photon-triggered quantum chaos,” arXiv:1905.04728 (2019).

29. C. Benedetti, F. Galve, A. Mandarino, M. G. Paris, and R. Zambrini, “Minimal model for spontaneous quantum synchronization,” Phys. Rev. A 94(5), 052118 (2016). [CrossRef]  

30. L. Du, C.-H. Fan, H.-X. Zhang, and J.-H. Wu, “Synchronization enhancement of indirectly coupled oscillators via periodic modulation in an optomechanical system,” Sci. Rep. 7(1), 15834 (2017). [CrossRef]  

31. A. Roulet and C. Bruder, “Quantum synchronization and entanglement generation,” Phys. Rev. Lett. 121(6), 063601 (2018). [CrossRef]  

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

33. 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]  

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

35. X.-B. Yan, C.-L. Cui, K.-H. Gu, X.-D. Tian, C.-B. Fu, and J.-H. Wu, “Coherent perfect absorption, transmission, and synthesis in a double-cavity optomechanical system,” Opt. Express 22(5), 4886–4895 (2014). [CrossRef]  

36. Y. Tokura and N. Nagaosa, “Nonreciprocal responses from non-centrosymmetric quantum materials,” Nat. Commun. 9(1), 3740 (2018). [CrossRef]  

37. L. Du, Y.-T. Chen, J.-H. Wu, and Y. Li, “Nonreciprocal quantum interference and coherent photon routing in a three-port optomechanical system,” arXiv:1909.07753 (2019).

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

39. 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]  

40. Y. Jiang, S. Maayani, T. Carmon, F. Nori, and H. Jing, “Nonreciprocal phonon laser,” Phys. Rev. Appl. 10(6), 064037 (2018). [CrossRef]  

41. I. M. Mirza, W. Ge, and H. Jing, “Optical nonreciprocity and slow light in coupled spinning optomechanical resonators,” Opt. Express 27(18), 25515–25530 (2019). [CrossRef]  

42. A. Metelmann and A. A. Clerk, “Nonreciprocal photon transmission and amplification via reservoir engineering,” Phys. Rev. X 5(2), 021025 (2015). [CrossRef]  

43. K. Fang, J. Luo, A. Metelmann, M. H. Matheny, F. Marquardt, A. A. Clerk, and O. Painter, “Generalized non-reciprocity in an optomechanical circuit via synthetic magnetism and reservoir engineering,” Nat. Phys. 13(5), 465–471 (2017). [CrossRef]  

44. S. Manipatruni, J. T. Robinson, and M. Lipson, “Optical nonreciprocity in optomechanical structures,” Phys. Rev. Lett. 102(21), 213903 (2009). [CrossRef]  

45. D. Malz, L. D. Tóth, N. R. Bernier, A. K. Feofanov, T. J. Kippenberg, and A. Nunnenkamp, “Quantum-limited directional amplifiers with optomechanics,” Phys. Rev. Lett. 120(2), 023601 (2018). [CrossRef]  

46. C. Jiang, L. N. Song, and Y. Li, “Directional phase-sensitive amplifier between microwave and optical photons,” Phys. Rev. A 99(2), 023823 (2019). [CrossRef]  

47. F. Ruesink, M.-A. Miri, A. Alu, and E. Verhagen, “Nonreciprocity and magnetic-free isolation based on optomechanical interactions,” Nat. Commun. 7(1), 13662 (2016). [CrossRef]  

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

49. F. Ruesink, J. P. Mathew, M.-A. Miri, A. Alù, and E. Verhagen, “Optical circulation in a multimode optomechanical resonator,” Nat. Commun. 9(1), 1798 (2018). [CrossRef]  

50. H. Xu, L. Jiang, A. Clerk, and J. Harris, “Nonreciprocal control and cooling of phonon modes in an optomechanical system,” Nature 568(7750), 65–69 (2019). [CrossRef]  

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

52. 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]  

53. L. Du, Y.-M. Liu, B. Jiang, and Y. Zhang, “All-optical photon switching, router and amplifier using a passive-active optomechanical system,” EPL 122(2), 24001 (2018). [CrossRef]  

54. B. He, L. Yang, X. Jiang, and M. Xiao, “Transmission nonreciprocity in a mutually coupled circulating structure,” Phys. Rev. Lett. 120(20), 203904 (2018). [CrossRef]  

55. B. Chen, C. Jiang, and K.-D. Zhu, “Slow light in a cavity optomechanical system with a bose-einstein condensate,” Phys. Rev. A 83(5), 055803 (2011). [CrossRef]  

56. M. J. Akram, M. M. Khan, and F. Saif, “Tunable fast and slow light in a hybrid optomechanical system,” Phys. Rev. A 92(2), 023846 (2015). [CrossRef]  

57. Z. Wu, R.-H. Luo, J.-Q. Zhang, Y.-H. Wang, W. Yang, and M. Feng, “Force-induced transparency and conversion between slow and fast light in optomechanics,” Phys. Rev. A 96(3), 033832 (2017). [CrossRef]  

58. S. Shahidani, M. H. Naderi, and M. Soltanolkotabi, “Control and manipulation of electromagnetically induced transparency in a nonlinear optomechanical system with two movable mirrors,” Phys. Rev. A 88(5), 053813 (2013). [CrossRef]  

59. B. P. Hou, L. F. Wei, and S. J. Wang, “Optomechanically induced transparency and absorption in hybridized optomechanical systems,” Phys. Rev. A 92(3), 033829 (2015). [CrossRef]  

60. I. Moaddel Haghighi, N. Malossi, R. Natali, G. Di Giuseppe, and D. Vitali, “Sensitivity-bandwidth limit in a multimode optoelectromechanical transducer,” Phys. Rev. Appl. 9(3), 034031 (2018). [CrossRef]  

61. M. Schmidt, S. Kessler, V. Peano, O. Painter, and F. Marquardt, “Optomechanical creation of magnetic fields for photons on a lattice,” Optica 2(7), 635–641 (2015). [CrossRef]  

62. M. O. Scully and M. S. Zubairy, “Quantum langevin equations,” Quantum Opt. (Cambridge Univ. Press., New York, 1997).

63. C. Gardiner and M. Collett, “Input and output in damped quantum systems: Quantum stochastic differential equations and the master equation,” Phys. Rev. A 31(6), 3761–3774 (1985). [CrossRef]  

64. M. Cai, O. Painter, and K. J. Vahala, “Observation of critical coupling in a fiber taper to a silica-microsphere whispering-gallery mode system,” Phys. Rev. Lett. 85(1), 74–77 (2000). [CrossRef]  

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

66. X. Z. Zhang, L. Tian, and Y. Li, “Optomechanical transistor with mechanical gain,” Phys. Rev. A 97(4), 043818 (2018). [CrossRef]  

67. E. X. DeJesus and C. Kaufman, “Routh-hurwitz criterion in the examination of eigenvalues of a system of nonlinear ordinary differential equations,” Phys. Rev. A 35(12), 5288–5290 (1987). [CrossRef]  

68. X. Xu and J. M. Taylor, “Squeezing in a coupled two-mode optomechanical system for force sensing below the standard quantum limit,” Phys. Rev. A 90(4), 043848 (2014). [CrossRef]  

69. S. Huang and G. S. Agarwal, “Robust force sensing for a free particle in a dissipative optomechanical system with a parametric amplifier,” Phys. Rev. A 95(2), 023844 (2017). [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 (6)

Fig. 1.
Fig. 1. (a) Schematic diagram of an optomechanical system composed of three indirectly coupled cavities and a mechanical resonator. (b) Schematic illustration of this system composed of three Fabry-Pérot-type resonators with a common movable mirror driven by a mechanical field. (c) Transition relation diagram of internal states in this multi-mode system where $n_{ak}$ and $n_{b}$ are the excitation numbers of the three cavity modes and the mechanical mode, respectively.
Fig. 2.
Fig. 2. (a) Stability diagram of the system on the plane of coupling strengths $G_{1}$ and $G_{3}$ for $\phi =0$ (The diagram is the same as those with $\phi$ of any value.). (b) Stability diagram of the system on the plane of phase difference $\phi$ and the coupling strength $G_{1}$ with $n=2$ . The other parameter is $\gamma _{m}=\kappa /100$ . The yellow regions are stable and the green regions are unstable.
Fig. 3.
Fig. 3. Logarithms of transmissivities (a) $T_{12}$ and (b) $T_{21}$ versus detuning $\delta$ and coupling strength $G_{3}$ for cavity $a_{3}$ on blue sideband with $y=2$ . Logarithms of the transmissivities $T_{12}$ and $T_{21}$ with $G_{3}=2.8\kappa$ and $y=2$ (c) for $a_{3}$ on blue sideband; (d) for $a_{3}$ on red sideband. The transmissivity $T_{12}$ versus $\delta$ with (e) $G_{3}=0$ and (f) $G_3=2.8\kappa$ for cavity $a_{3}$ on blue sideband. Other parameters are $\phi =\pi$ , $\beta =\pi /2$ , $\gamma _{m}=\kappa /100$ and $G=2\kappa$ .
Fig. 4.
Fig. 4. (a) Logarithms of the transmissivities $T_{21}$ and $T_{12}$ and (b) isolation ratio $I$ versus relative ratio $y$ with $\delta =0$ . Other parameters are the same as those in Fig. 3.
Fig. 5.
Fig. 5. Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\delta$ for $\phi =\pi /2$ and (a) $\beta =0$ ; (b) $\beta =\pi /2$ . Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\beta$ for $\delta =0$ and (c) $\phi =\pi /2$ ; (d) $\phi =\pi$ . Logarithms of transmissivities $T_{21}$ and $T_{12}$ verse $\phi$ for $\delta =0$ and (e) $\beta =3\pi /2$ ; (f) $\beta =\pi$ . (g) Isolation ratio $I$ verse $\beta$ for $\phi =\pi /2$ and $\delta =0$ . (h) Transmissivities $T_{21}$ and $T_{12}$ verse $\phi$ for $\beta =\pi /2$ . Other parameters are the same as those in Fig. 3.
Fig. 6.
Fig. 6. (a) Group delay $\tau _{12}$ versus coupling strength $G_{3}$ for $\phi =\pi /2$ , $\beta =\pi /2$ and $\delta =0$ . (b) Group delays $\tau _{12}$ and $\tau _{21}$ versus detuning $\delta$ for $\phi =\pi$ and $\beta =0$ . $\tau _{12(21)}=0$ is the transition point of slow and fast lights. Other parameters are the same as those in Fig. 3.

Equations (19)

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

H = H 0 + H I + H d .
H 0 = k = 1 3 ω k a k a k + ω m b b ;
H I = g 1 a 1 a 1 ( b + b ) + g 2 a 2 a 2 ( b + b ) + g 3 a 3 a 3 ( b + b ) ;
H d = k = 1 3 ε c k ( i a k e i ω c k t e i θ k + H . c . ) + ε p ( i a 1 e i ω p t + H . c . ) + ε b ( i b e i ω b t + H . c . ) ,
a 1 ˙ = ( κ + i Δ 1 ) a 1 + i g 1 a 1 ( b + b ) + ε c 1 e i θ 1 + ε p e i σ t + 2 κ a 1 i n , a 2 ˙ = ( κ + i Δ 2 ) a 2 i g 2 a 2 ( b + b ) + ε c 2 e i θ 2 + 2 κ a 2 i n , a 3 ˙ = ( κ + i Δ 3 ) a 3 i g 3 a 3 ( b + b ) + ε c 3 e i θ 3 + 2 κ a 3 i n , b ˙ = ( γ m + i ω m ) b + i ( g 1 a 1 a 1 g 2 a 2 a 2 g 3 a 3 a 3 ) + ε b e i ω b t + 2 γ m b i n ,
a 1 = ε c 1 e i θ 1 κ + i Δ 1 , a 2 = ε c 2 e i θ 2 κ + i Δ 2 , a 3 = ε c 3 e i θ 3 κ + i Δ 3 , b = i ( g 1 | α 1 | 2 g 2 | α 2 | 2 g 3 | α 3 | 2 ) γ m + i ω m ,
δ a 1 ˙ = ( κ + i Δ 1 ) δ a 1 + i G 1 ( δ b + δ b ) + ε p e i σ t + 2 κ a 1 i n , δ a 2 ˙ = ( κ + i Δ 2 ) δ a 2 i G 2 ( δ b + δ b ) + 2 κ a 2 i n , δ a 3 ˙ = ( κ + i Δ 3 ) δ a 3 i G 3 ( δ b + δ b ) + 2 κ a 3 i n , δ b ˙ = ( γ m + i ω m ) δ b + i ( G 1 δ a 1 + G 1 δ a 1 G 2 δ a 2 G 2 δ a 2 G 3 δ a 3 G 3 δ a 3 ) + ε b e i ω b t + 2 γ m b i n ,
δ a 1 ˙ = κ δ a 1 + i G 1 δ b + ε p e i δ t + 2 κ a 1 i n , δ a 2 ˙ = κ δ a 2 i G 2 δ b + 2 κ a 2 i n , δ a 3 ˙ = κ δ a 3 + i G 3 δ b + 2 κ a 3 i n , δ b ˙ = γ m δ b + i G 1 δ a 1 i G 2 δ a 2 i G 3 δ a 3 + ε b e i δ t + 2 γ m b i n ,
δ a 1 ˙ = ( κ i δ ) δ a 1 + i G 1 δ b + ε p , δ a 2 ˙ = ( κ i δ ) δ a 2 i G 2 δ b , δ a 3 ˙ = ( κ i δ ) δ a 3 + i G 3 δ b , δ b ˙ = ( γ m i δ ) δ b + i G 1 δ a 1 i G 2 δ a 2 i G 3 δ a 3 + ε b .
δ a 1 = | G | 2 ε p + i G ε b ( κ i δ ) N + ε p κ i δ , δ a 2 = G 2 e i ϕ ε p i G e i ϕ ε b ( κ i δ ) N
t 21 = 2 κ [ G 2 e i ϕ i G y e i ( ϕ + β ) ( κ i δ ) ] N
t 12 = 2 κ [ G 2 e i ϕ i G y e i β ( κ i δ ) ] N .
τ ˙ = M τ + ν + μ ,
M = ( κ 0 0 i G 1 0 κ 0 i G 2 0 0 κ i G 3 i G 1 i G 2 i G 3 γ m   ) .
( γ m i δ + | G 1 | 2 + | G 2 | 2 | G 3 | 2 κ i δ ) δ b = ( i G 1 κ i δ + y e i β ) ε p ,
Γ b = γ m + κ ( 2 | G | 2 | G 3 | 2 ) κ 2 + δ 2 ,
Γ r = γ m + κ ( 2 | G | 2 + | G 3 | 2 ) κ 2 + δ 2 .
I = T 12 T 21 T 12 + T 21
τ 12 ( 21 ) = d Θ 12 ( 21 ) d ω p ,
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.