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

Optical noise-resistant nonreciprocal phonon blockade in a spinning optomechanical resonator

Open Access Open Access

Abstract

A scheme of nonreciprocal conventional phonon blockade (PB) is proposed in a spinning optomechanical resonator coupled with a two-level atom. The coherent coupling between the atom and breathing mode is mediated by the optical mode with a large detuning. Due to the Fizeau shift caused by the spinning resonator, the PB can be implemented in a nonreciprocal way. Specifically, when the spinning resonator is driven from one direction, the single-phonon (1PB) and two-phonon blockade (2PB) can be achieved by adjusting both the amplitude and frequency of the mechanical drive field, while phonon-induced tunneling (PIT) occurs when the spinning resonator is driven from the opposite direction. The PB effects are insensitive to cavity decay because of the adiabatic elimination of the optical mode, thus making the scheme more robust to the optical noise and still feasible even in a low-Q cavity. Our scheme provides a flexible method for engineering a unidirectional phonon source with external control, which is expected to be used as a chiral quantum device in quantum computing networks.

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

1. Introduction

As well known, analogy to photon blockade [18], Coulomb blockade [9], and magnon blockade [10,11], PB is a pure quantum phenomenon which refers to the anti-bunching effect of the phonon in a system. It provides the possibility to develop novel phonon devices at the single-phonon level [12]. So far, there are two kinds of PB mechanisms: one is based on the destructive interference between different excitation paths, i.e., unconventional PB, which has been proposed in nanomechanical resonator [13,14] and optomechanical systems [1517]; another is based on the anharmonicity of the energy levels, known as conventional PB, which has been realized by coupling a mechanical oscillator to a qubit [18,19] or a two-level defect [20]. On the contrary, PIT is a completely different phenomenon from PB, that is, absorption of the first phonon is conducive to the absorption of subsequent phonons, and it is often regarded as a potential tool for generating specific multi-phonon states. Both of these two physical effects are important for developing the phonon devices.

Nonreciprocal devices have garnered significant attention due to their ability to enable unidirectional light transmission, which is crucial for various information processing applications [2124]. It has been demonstrated in various physical systems such as atomic gases [25,26], non-Hermitian optics [27,28], the nonlinear cavity [2932]. Different from the classical nonreciprocity, quantum nonreciprocal devices are used to realize nonreciprocal quantum effects based on quantum regime [3335]. Very recently, an optical diode with $99.6\%$ isolation has been observed by using a spinning resonator [36]. By spinning a shaped-deformed resonator, a variety of works have been studied such as nonreciprocal entanglement [37,38], nonreciprocal phonon laser [39], nonreciprocal optical soliton [40], nonreciprocal slow light [41,42], and enhanced sensing [43,44]. Nonreciprocal photon blockade and magnon blockade have been proposed in a spinning Kerr cavity [45], a quadratic optomechanical coupling system [46] and a magnon-based hybrid system [47], respectively. Such devices can be applied to the quantum control of light in chiral quantum computing and topological quantum techniques. Xue et al. have proposed to achieve nonreciprocal photon blockade in a spinning resonator in which the Fizeau shifts can be flexibly adjusted [48]. Li et al. have investigated the unconventional photon blockade in a spinning optomechanical resonator [49]. However, the nonreciprocal phonon blockade effect induced by a rotating optomechanical resonator is rarely studied. The optomechanical system is usually composed of the optical cavity coupled with the mechanical oscillator through radiation pressure interaction, which can be used to change the properties of the optical field or control the behavior of the mechanical oscillator. It has been found that many unique and interesting phenomena can occur in the optomechanical resonator, such as enhancement of nonlinearity [50,51], phonon laser [5254], super-resolution mass sensing [55], etc. In addition, improving the anti-noise ability of quantum systems is always a concern in development of quantum devices and quantum information processing. Some strategies have been proposed to resist optical noise or avoid backscattering, e.g., by introducing additional scattering [56], synthetic gague fields [57], phased-controlled asymmetric optomechanical system [58], and using rotating cavity [37]. These anti-noise quantum technologies can improve the stability and reliability of quantum systems. However, it should be noted that noise-resistant quantum devices do not completely eliminate noise, but minimize the effect of noise through effective control and suppression. Therefore, the exploration of noise-resistant quantum nonreciprocal devices is worthy of further study.

Inspired by these contributes, here we study nonreciprocal conventional PB in a spinning optomechanical resonator assisted by a two-level atom by considering the radiation pressure interaction between the breathing mode and optical mode. The analytical energy spectrum of the atom-phonon subsystem and the optimal condition of PB are obtained by adiabatically eliminating the optical mode with large detuning. We show that PB effects and PIT can be realized in the spinning resonator in a nonreciprocal manner. Furthermore, either the nonreciprocal 1PB or 2PB can be achieved by varying both the amplitude and frequency of the driving field applied to the breathing mode. Since the optical mode is adiabatically eliminated, the PB is insensitive to the cavity decay, which enables the scheme more robust to optical noise. There is no strict restriction on the quality of the resonator, that is, the scheme is still feasible even in a low-Q cavity. It can be used as a robust unidirectional phonon source in one-way quantum computing networks.

The remainder of the paper is organized as follows. In Sec. 2, we describe the physical model and derive the effective Hamiltonian of the system. In Sec.3, the optimal condition of PB is obtained by calculating the eigenenergy spectrum. In Sec.4, we investigate the nonreciprocal PB effect including 1PB and 2PB with the equal-time second- and third-order correlation functions by using numerical simulation of master equation. In Sec.5, we discuss the validity of present scheme and the influence of thermal phonons on PB, and conclude the results finally.

2. Model and effective Hamiltonian

The hybrid system considered here consists of a spinning toroidal resonator supporting a breathing mode, and a two-level atom with frequency $\omega _a$ is arranged near the resonator so that coupling with the optical modes, as shown in Fig. 1. In the spinning resonator, the clockwise (CW) or counterclockwise (CCW) optical mode undergoes different refractive indices. Additionally, a weak optomechanical coupling between the optical mode and the breathing mode is considered. For a resonator spinning at an angular velocity $\Omega$, the optical mode experiences a Fizeau shift $\Delta _F$, leading to the optical resonance frequency of the optomechanical resonator $\omega _c$ changed, that is, $\omega _c\rightarrow \omega _c+\Delta _F$ [59], with

$$\Delta_F={\pm}\frac{nr\Omega\omega_c}{c}\left(1-\frac{1}{n^2}-\frac{1}{n}\frac{dn}{d\lambda}\right),$$
where $n$ is the refractive index, $r$ is the resonator radius, and $c (\lambda )$ is the speed (wavelength) of the light in vacuum. Usually, the dispersion term $dn/d\lambda$ which denotes the relativistic origin of the Sagnac-Fizeau effect is relatively small (up to $1\%$) [36], so that it can be ignored safely. Here, $\Delta _F>0$ ($\Delta _F<0$) means that the driving light propagates against (along) the spinning direction of the resonator, i.e., the frequencies of CW and CCW mode are $\omega _c\pm |\Delta _F|$, respectively.

 figure: Fig. 1.

Fig. 1. A spinning optomechanical resonator coupled to a two-level atom. Spinning the resonator results in different Fizeau drag $\Delta _F$ for engineering the clockwise or counterclockwise optical mode undergoing different refractive indices with an angular velocity $\Omega$. (a) denotes driving the resonator from left side ($\Delta _F>0$), and (b) from right side ($\Delta _F<0$).

Download Full Size | PDF

In a rotating frame at the driving laser frequency $\omega _{L}$, the system Hamiltonian can be described as ($\hbar =1$)

$$\begin{aligned} H & =(\Delta_c+\Delta_F)a^{\dagger}a+\Delta_a\sigma_{+}\sigma_{-}+\omega_mb^{\dagger}b+J(a^{\dagger}\sigma_{-}+a\sigma_{+})\cr\cr & +ga^{\dagger}a(b^{\dagger}+b)+\varepsilon(a^{\dagger}+a), \end{aligned}$$
where $a (a^{\dagger })$ and $b (b^{\dagger })$ represent the annihilation (creation) operator for cavity and mechanical mode, respectively. $\sigma _{+}=\vert e\rangle \langle g\vert$ and $\sigma _{-}=\vert g\rangle \langle e\vert$ denote the raising and lowering operator of the two-level atom. $\Delta _i=\omega _i-\omega _{L}$ $(i=a,c)$ denotes the detuning of the two-level atom and the cavity to the driving laser and $\omega _m$ is the frequency of the mechanical mode. $J$ and $g$ denote the cavity-atom coupling strength and the optomechanical coupling strength, respectively. The parameter $\varepsilon$ is the driving amplitude of the cavity. By means of the standard linearization procedure of the cavity optomechanics, the operator $o$ is split into an average amplitude and a fluctuation term, i.e., $o\rightarrow \bar {o}+o$, thus the Hamiltonian in Eq. (2) can be written as
$$\begin{aligned} H_{1} & =(\Delta_c+\Delta_F)a^{\dagger}a+\Delta_a\sigma_{+}\sigma_{-}+\omega_mb^{\dagger}b+J(a^{\dagger}\sigma_{-}+a\sigma_{+})\cr\cr & +g\bar{a}(a^{\dagger}+a)(b^{\dagger}+b), \end{aligned}$$
where the small second-order term $ga^{\dagger }a(b^{\dagger }+b)$ has been omitted in comparison to $g\bar {a}(a^{\dagger }+a)(b^{\dagger }+b)$. In the case of $g\bar {a}\ll \omega _m$, the rapidly oscillating terms with high frequencies $\omega _m$ can be safely neglected under the rotating wave approximation. Therefore, the Hamiltonian of the system can be read approximately as
$$H_{2}=(\Delta_c+\Delta_F)a^{\dagger}a+\Delta_a\sigma_{+}\sigma_{-}+\omega_mb^{\dagger}b+J(a^{\dagger}\sigma_{-}+a\sigma_{+})+G(a^{\dagger}b+ab^{\dagger}),$$
where $G=g\bar {a}$ is the effective light-mechanical coupling enhanced by the cavity. By using the Fröhlich-Nakajima transformation [60,61], the cavity mode can be decoupled from the system, then the effective Hamiltonian of phonon-atom subsystem mediated by the cavity is obtained as
$$H_{3}=\left(\Delta_a+\dfrac{J^2}{\delta+\Delta_{F}}\right)\sigma_{+}\sigma_{-}+\left(\omega_m+\dfrac{G^2}{\delta+\Delta_{F}}\right)b^{\dagger}b+\left(\dfrac{JG}{\delta+\Delta_{F}}\right)(b^{\dagger}\sigma_{-}+b\sigma_{+}),$$
where $\delta _a=\Delta _a-\Delta _c$, $\delta _b=\omega _m-\Delta _c$, and $\Delta _a=\omega _m$ has been assumed so that $\delta _a=\delta _b=\delta$ (see Appendix A for derivation). Now using a weak drive field with frequency $\omega _p$ and amplitude $E$ to drive the mechanical mode, in the rotating frame defined by $U=\exp \left [-i\omega _pt(\sigma _{+}\sigma _{-}+b^{\dagger }b)\right ]$, the effective Hamiltonian can be written as
$$H_{\rm{eff}}=\left(-\Delta+\dfrac{J^2}{\delta+\Delta_{F}}\right)\sigma_{+}\sigma_{-}+\left(-\Delta+\dfrac{G^2}{\delta+\Delta_{F}}\right)b^{\dagger}b+\left(\dfrac{JG}{\delta+\Delta_{F}}\right)(b^{\dagger}\sigma_{-}+b\sigma_{+})+E(b^{\dagger}+b),$$
where $\Delta =\omega _p-\Delta _a=\omega _p-\omega _m$ are the effective detuning of the two-level atom and the mechanical mode, respectively. The effective Hamiltonian describes a Jaynes-Cumming type interaction between a phonon mode and a two-level atom.

3. Optimal condition of 1PB

In the subspace spanned by base vectors $\{|m,g\rangle,|m-1,e\rangle \}$, where $|m\rangle$ $(m=1,2,3,\ldots )$ represents Fock state of the phonon and $|g\rangle$ ($|e\rangle$) the ground (excited) state of the two-level atom, the effective Hamiltonian without involvement of the weak driving term can be written as a matrix form

$$H=\left(\begin{array}{cc} m\left(-\Delta+\dfrac{G^2}{\delta+\Delta_{F}}\right) ~~~ & ~~~ \dfrac{JG}{\delta+\Delta_{F}}\sqrt{m}\\ \dfrac{JG}{\delta+\Delta_{F}}\sqrt{m} ~~~ & ~~~ m\left(-\Delta+\dfrac{J^2}{\delta+\Delta_{F}}\right)\\ \end{array}\right).$$

The energy eigenvalues are calculated as

$$E_{m,\pm}=\frac{m[J^2+G^2-2\Delta(\delta+\Delta_{F})]\pm \sqrt{m}K}{2(\delta+\Delta_{F})},$$
corresponding to the unnormalized eigenstates
$$|m,\pm\rangle=\frac{\sqrt{m}(G^2-J^2)\pm K}{2JG}|m,g\rangle+|m-1,e\rangle,$$
with $K=\sqrt {4J^2G^2+m(G^2-J^2)^2}$. We can see from Eq. (8) that the eigenenergy is a nonlinear function of the phonon number $m$, and the anharmonicity of energy spectrum will disappear without the cavity-atom coupling (i.e., $J=0$) in this optomechanical system. The anharmonicity of the eigenenergy spectrum is the physical origin of the PB effect in the system.

In order to obtain the optimal parametric condition of 1PB, we truncate the largest phonon excitation number to $m=2$, as shown in the Fig. 2. Assuming that the system is initially in its ground state $|0,g\rangle$, we only need to consider the transitions involving the states $|0,g\rangle$, $|1,\pm \rangle$ and $|2,\pm \rangle$ due to the weak driving on mechanical mode. If the driving light resonantly excites the transition $|0,g\rangle \leftrightarrow |1,\pm \rangle$, the condition $\omega _p=E_{1,\pm }$ should be satisfied. Then we get the optimal condition for 1PB as

$$J^2={-}G^2+\Delta(\delta+\Delta_{F}).$$

 figure: Fig. 2.

Fig. 2. The energy spectrum of the phonon-atom subsystem in the few-phonon subspace when the angular velocity $\Omega =0$. $|m,g(e)\rangle$ represents the state with $m$ phonons and the atom in ground (excited) state. The green and purple arrows represent the frequency of driving laser.

Download Full Size | PDF

With this optimal condition in Eq. (10), the driving field with frequency $\omega _p$ will excite the system to a single-phonon state $|1,\pm \rangle$, but cannot further excite the system to higher phonon state due to the large detuning. In other words, once a phonon is excited in the system, it will prevent the second phonon with the same frequency from entering the hybrid system. Therefore, the strong 1PB can be implemented. Moreover, the optimal condition in Eq. (10) is dependent on Fizeau shift, which indicates the spinning angular velocity $\Omega$ can be used for tuning the PB effect.

4. Nonreciprocal PB effects

To investigate the effect of thermal noise on the quantum statistics of the phononic field, the quantum master equation of the system reads [62,63]

$$\begin{aligned} \dot\rho= & -i\left[H_{\rm{eff}},\rho\right]+\frac{\gamma_a}{2}\left(2\sigma_{-}\rho \sigma_{+}-\sigma_{+}\sigma_{-}\rho-\rho \sigma_{+}\sigma_{-}\right)\cr\cr & +\frac{\gamma_{m}}{2}(n_{th}+1)\left(2b\rho b^{\dagger}-b^{\dagger}b\rho-\rho b^{\dagger}b\right)\cr\cr & +\frac{\gamma_{m}}{2}n_{th}\left(2b^{\dagger}\rho b-bb^{\dagger}\rho-\rho bb^{\dagger}\right), \end{aligned}$$
where $\gamma _{a}$ is the decay rate of the atom, and the damping rate of the mechanical mode is $\gamma _{m}=\omega _m/Q_M$ with $Q_M$ the quality factor of the mechanical mode. $\rho$ is the dynamical density matrix of the system, $n_{th}$=exp$[\hbar \omega _m/(k_{B}T)-1]^{-1}$ is the thermal mean phonon number with the ambient temperature $T$ and Boltzmann constant $k_{B}$. In the following numerical calculations, we take $\gamma _a=\gamma _{m}=\gamma$. The statistical distribution of phonon in steady-state can be described by the equal-time second-order correlation function
$$g^{(2)}{(0)}=\frac{Tr\left(\rho b^{\dagger}b^{\dagger}bb\right)}{\left[Tr\left(\rho b^{\dagger}b\right)\right]^{2}}=\frac{\langle b^{\dagger}b^{\dagger}bb\rangle}{\langle b^{\dagger}b\rangle^{2}}.$$

In general, the value of $g^{(2)}{(0)}$ characterizes the probability for detecting two phonons simultaneously. The case of $g^{(2)}{(0)}>1$ implies a super-Poissonian statistics of the phonons, which is also referred to as bunching effect. The case of $g^{(2)}{(0)}<1$ implies the sub-Poissonian statistics of the phonons, which is referred to as phonon antibunching effect, i.e., detecting the phonons one by one. Especially, $g^{(2)}{(0)}=0$ denotes a perfect PB effect.

4.1 Nonreciprocal 1PB effect

We firstly study the 1PB effect for a stationary resonator (i.e., $\Omega =0$) by numerically calculating the equal-time second-order correlation function and average phonon number of the mechanical mode. Firstly, the logarithmic second-order correlation function $\lg [g^{(2)}(0)]$ and the average phonon number $N_{\rm b}$ are plotted as the functions of ($J/\kappa$, $\Delta /\kappa$) in Fig. 3(a) and 3(c), and as the functions of ($G/\kappa$, $\Delta /\kappa$) in Fig. 3(b) and 3(d), respectively, for absolute temperature $T$=0K. Here, the red contour delineates the case of second-order correlation function $g^{(2)}(0)=1$, which is the boundary between the super-Poissonian light ($\lg g^{(2)}(0)>0$) and sub-Poissonian light ($\lg g^{(2)}(0)<0$). The results of numerical simulation show that when the logarithmic second-order correlation function is less than 0 in Fig. 3(a) and 3(b), the average phonon number has the maximum value in Fig. 3(c) and 3(d), the parameters fully conform to the optimal condition in Eq. (10). However, there is no PB effect for $J=0$ or $G=0$ in that the anharmonicity of eigenenergy spectrum disappears, which can be seen from Eqs. (7)–(8).

 figure: Fig. 3.

Fig. 3. Logarithmic second-order correlation function $\lg {g^{(2)}(0)}$ in (a) and the mean phonon number $N_{b}$ in (c) as functions of $J/\kappa$ and $\Delta /\kappa$ for $G/\kappa =10$. Logarithmic second-order correlation function $\lg {g^{(2)}(0)}$ in (b) and mean phonon number $N_{b}$ in (d) as functions of $G/\kappa$ and $\Delta /\kappa$ for $J/\kappa =10$. The white dotted line denotes the optimal conditions of the conventional PB in Eq. (10). The red contour denotes the case of second-order correlation function $g^{(2)}(0)=1$. The other parameters are taken as $\Omega =0$, $E/\kappa =0.01$, $\delta /\kappa =50$, $n_{th}=0$, $\gamma /\kappa =0.03$.

Download Full Size | PDF

Next, we discuss nonreciprocal PB effect for a spinning optomechanical resonator by taking the experimentally feasible parameters [64]: $Q=5\times 10^{9}$, $\lambda =1550nm$, $r=0.3mm$, $n=1.4$. Moreover, spinning objects have achieved significantly higher velocities, reaching the GHz range [65,66]; such systems also can be used to research the nonreciprocal PB through quadratic optomechanical coupling [67,68] and Kerr-like optomechanical interactions [69]. In order to observe the nonreciprocal PB effect more clearly, we assume that 1PB and two-phonon tunneling occur when the driving laser input from the left and right side, respectively. Considering this limitation and the eigenenergy spectrum in Eq. (8), an optimal Fizeau shift $|\Delta _F/\kappa |=3.95043$ can be calculated by the method in Ref. [48], and the corresponding detuning $\Delta /\kappa =3.70711$ can be derived. In Fig. 4(a), the second-order correlation function $g^{(2)}{(0)}$ is plotted versus the detuning $\Delta /\kappa$ for $\Delta _F>0$ and $\Delta _F<0$, respectively. It has been shown that by taking $\Delta /\kappa =3.70711$, a dip “a” with a negative value appears when the driving laser is input from the left side (the blue solid curve), implying that the 1PB effect can be observed. While a peak “b” with a positive value appears when the driving laser is input from the right side (the red dashed curve), i.e., the PIT occurs. This is agree with our expectation. This quantum nonreciprocity can be found with a difference of up to 5 orders of magnitude between two second-order correlation functions in opposite directions.

 figure: Fig. 4.

Fig. 4. (a) Second-order correlation function $g^{(2)}{(0)}$ versus the detuning $\Delta /\kappa$ for different driving directions of the cavity. The nonreciprocal 1PB occurs in solved detuning region at $\Delta /\kappa =3.70711$ and $\Delta _{F}/\kappa =\pm 3.95043$. (b) The energy levels when driving the system from left and right side is nonequidistant for $\Delta /\kappa =3.70711$. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

Interestingly, the enhancement of PB effect can be observed for $\Delta _F<0$ in Fig. 4(a) by comparing to the cases of $\Delta _F>0$ and $\Delta _F=0$. This can be explained by the fact that for $J=G$, the excitation $\vert 1,+\rangle _{L} \rightarrow |2,+\rangle _{L}$ is far from resonance since the detuning $\dfrac {(2-\sqrt {2})JG}{\delta +\Delta _{F}}\gg \{\kappa,\gamma \}$ calculated from Eq. (8). Consequently, a larger detuning is induced for $\Delta _{F}<0$, resulting in a lower occupancy of the two-phonon state $\vert 2,+\rangle _{L}$. As a result, PB effect is enhanced. This indicates that a spinning cavity may be used to enhance PB effect.

The energy eigenstates are shown in Fig. 4(b), where $|m,+\rangle _{L(R)}$ denotes the related state of $m$-phonon when the optomechanical resonator is driven from the left (right) side. We can see that when the driving frequency on the phonon is resonantly coupled to the transition $|0,g\rangle _{L} \rightarrow |1,+\rangle _{L}$, the transition $|1,+\rangle _{L} \rightarrow |2,+\rangle _{L}$ will be suppressed for the large detuning. On the other hand, there will be a two-phonon resonance $|0,g\rangle _{R} \rightarrow |2,+\rangle _{R}$ under the same laser frequency, thus a process of PIT can be induced when the resonator is driven from the right side. Therefore, the nonreciprocal conventional 1PB in such spinning device is achieved.

4.2 Nonreciprocal 2PB effect

Now we further discuss the nonreciprocity of 2PB and PIT effect, which can be intuitively understood from the anharmonicity of energy structure as shown in Fig. 5. When the cavity is driven from the left side ($\Delta _{F}>0$), it is required that the light resonants with the $|0\rangle \rightarrow |2\rangle$ transition, but the transition $|2\rangle \rightarrow |3\rangle$ is suppressed, which features the 2PB effect; in contrast, the three-phonon resonance happens for the transition $|0\rangle \rightarrow |3\rangle$ when the cavity is driven from the right side ($\Delta _{F}<0$), which leads to three-phonon tunneling. This nonreciprocity can be understood by numerically analyzing the equal-time second- and three-order correlation functions of mechanical mode $g^{(2)}(0)$ and $g^{(3)}(0)$, i.e., [70]

$$g^{(2)}(0)>1 \quad \& \quad g^{(3)}(0)<1.$$

 figure: Fig. 5.

Fig. 5. The energy level spectrum of nonreciprocal 2PB. With enhanced driving strength $E/\kappa =0.2$ and the same Fizeau drag $\Delta _{F}/\kappa =\pm 3.95043$, 2PB occurs when the resonator is driven from left side ($\Delta _{F}>0$), while three-phonon induced PIT happens when the resonator is driven from right side ($\Delta _{F}<0$).

Download Full Size | PDF

In the meantime, the phonon-number probability ${P}(m)=\langle m|\rho |m\rangle$ can be obtained from the steady-state solution $\rho$ of the master equation in Eq. (11). A relative deviation of a given phonon-number distribution from the corresponding Poissonian distribution $\mathcal {P}(m)$ is used for describing the statistical property of the phonon distribution, i.e.,

$$[{P}(m)-\mathcal{P}(m)]/\mathcal{P}(m),$$
where $\mathcal {P}(m)=\dfrac {\overline {m}^m}{m!}e^{-\overline {m}}$ is the Poissonian distribution, and $\overline {m}$ is the average phonon number in the mechanical mode.

Under the limit of the condition in Eq. (13), the nonreciprocal 2PB effect can be achieved by increasing the driving strength $E$ and tuning driving frequency $\omega _p$ of the mechanical mode. The correlation function is depicted versus the detuning $\Delta /\kappa$ for $E/\kappa =0.2$ and $\Delta _{F}/\kappa =\pm 3.95043$ in the Fig. 6(a), where the blue curves represent $g^{(n)}_{L}(0)$ under the left driving ($\Delta _{F}>0$), and the red curves represent $g^{(n)}_{R}(0)$ under the right driving ($\Delta _{F}<0$) $(n=2,3)$. It can be clearly observed from the enlarged picture 6(b) that 2PB occurs in the area of light blue strip around detuning $\Delta /\kappa =3.25$ for the left driving, that is, the correlation functions $g^{(2)}(0)$ and $g^{(3)}(0)$ satisfy the criteria given in Eq. (13). Besides, the nonreciprocal 2PB can also be confirmed by comparing the phonon-number distribution ${P}(n)$ with the Poissonian distribution $\mathcal {P}(n)$. The Fig. 6(c) shows that two-phonon probability ${P}(2)$ is significantly enhanced while other-phonon probabilities are suppressed for left driving ($\Delta _{F}>0$), leading to 2PB. In contrast, the Fig. 6(d) shows that there is no 2PB occurs for the right driving ($\Delta _{F}<0$) but a three-phonon resonance, resulting in the three-phonon tunneling.

 figure: Fig. 6.

Fig. 6. (a) Second-order correlation function $g_{L,R}^{(2)}(0)$ (solid curves) and three-order correlation function $g_{L,R}^{(3)}(0)$ (dashed curves) versus the frequency detuning for different driving directions, and the nonreciprocal 2PB can be observed clearly by the enlarged figure in (b). Relative deviation of the phonon population from the Poisson distribution is shown in (c) when the nonreciprocal 2PB occurs and in (d) when the PIT occurs.

Download Full Size | PDF

Here, we should note that the energy level separation between $|1,-\rangle$ and $|1,+\rangle$ and that between $|2,-\rangle$ and $|2,+\rangle$ should be large enough to ensure the level anharmonicity. Here, the driving frequency of cavity can be adjusted to enhance cavity detuning and increase the energy level spacing. Thus ensures the transition $|1,+\rangle \rightarrow |2,+\rangle$ cannot be influenced by the other transitions $|1,+\rangle \rightarrow |2,-\rangle$, and $|1,-\rangle \rightarrow |2,\pm \rangle$.

5. Discussions and conclusions

To verify the validity of the theoretical scheme, the equal-time second-order correlation function $g^{(2)}(0)$ is simulated by using the linearized Hamiltonian $H_2$ and the effective Hamiltonian $H_{\rm {eff}}$, respectively. Considering the driving field with frequency $\omega _p$ of the mechanical mode, the Hamiltonian $H_2$ can be rewritten in a rotating frame defined by $U_{1}=\exp {[-i\omega _pt(a^{\dagger }a+b^{\dagger }b+\sigma _{+}\sigma _{-})]}$ as

$$H_{l}= \tilde{\Delta}_{c}a^{\dagger}a+\Delta \sigma_{+}\sigma_{-}+\Delta b^{\dagger}b+J(a^{\dagger}\sigma_{-}+a\sigma_{+})+G(a^{\dagger}b+ab^{\dagger})+E(b^{\dagger}+b),$$
where $\tilde {\Delta }_{c}=\Delta _{c}-\omega _p$, $\Delta =\Delta _a-\omega _p=\omega _m-\omega _p$. Considering the influence of cavity decay, the master equation can be written as
$$\begin{aligned} \dot\rho= & -i\left[H_{l},\rho\right]+\frac{\kappa}{2}\left(2a\rho a^{\dagger}-a^{\dagger}a\rho-\rho a^{\dagger}a\right)\cr\cr & +\frac{\gamma}{2}\left(2\sigma_{-}\rho \sigma_{+}-\sigma_{+}\sigma_{-}\rho-\rho \sigma_{+}\sigma_{-}\right)+\frac{\gamma}{2}(n_{th}+1)\left(2b\rho b^{\dagger}-b^{\dagger}b\rho-\rho b^{\dagger}b\right)\cr\cr & +\frac{\gamma}{2}n_{th}\left(2b^{\dagger}\rho b-bb^{\dagger}\rho-\rho bb^{\dagger}\right), \end{aligned}$$
where $\kappa$ represents the decay rate of the cavity. We can see from the Fig. 7 that the second-order correlation functions governed by $H_l$ (blue dashed line) and $H_{\rm {eff}}$ (red solid line) coincide well with each other even in a cavity with large dissipation (for example, the cavity decay rate is 100 times of the driving amplitude, and 33 times of the damping rates of the atom and the mechanical mode). Hence, the nonreciprocal PB effects are insensitive to the cavity decay due to the adiabatic elimination, which enables the scheme robust to optical noise. This is very advantageous in experiments because there is no strict restriction on the quality of the resonator.

 figure: Fig. 7.

Fig. 7. The numerical results of second-order correlation function $g^{(2)}(0)$ simulated by the linearized Hamiltonian $H_l$ and the effective Hamiltonian $H_{\rm {eff}}$ coincide well with each other. The parameters are taken as $\tilde {\Delta }_{c}/\kappa =50$, $J/\kappa =G/\kappa =10$, $E/\kappa =0.01$, $\gamma /\kappa$=0.03.

Download Full Size | PDF

Additionally, the effect of thermal phonons on PB is considered. The $g^{(2)}(0)$ is plotted as a function of $\Delta /\kappa$ for different thermal phonon numbers in Fig. 8. It can be seen that the thermal phonons has an adverse effect on PB with the temperature increasing. Therefore, in order to observe the PB effect, the influence of thermal phonons must be overcome, for instance, by lowering the ambient temperature or cooling the mechanical mode to its ground state.

 figure: Fig. 8.

Fig. 8. The second-order correlation function $g^{(2)}(0)$ as a function of the detuning $\Delta /\kappa$ for different thermal phonon numbers $n_{th}$. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

Finally, the current scheme of non-reciprocal PB is implemented by adjusting the frequency and amplitude of the mechanical driving field. The mechanical driving for a ring resonator can be realized by leveraging the experimental technique outlined in the Ref. [71]. In this study, researchers designed and fabricated a suspended ring resonator using piezoelectric material aluminum nitride (AlN), which simultaneously functions as an optical cavity for sensitive displacement readout. The mechanical mode can be driven by electrical signal.

In conclusion, we have presented a scheme to realize the nonreciprocal conventional PB in a spinning optomechanical resonator assisted by a two-level atom. In this scheme, by adjusting both the amplitude and frequency of the driving field applied to the mechanical mode, the 1PB and 2PB can be implemented in the nonreciprocal way. The validity of the scheme is confirmed by comparison between the results of numerical simulation governed, respectively, by the linearized and effective Hamiltonian. Since the optical mode is adiabatically eliminated, the PB effect is insensitive to the cavity decay, which makes the scheme more robust to optical noise, i.e., it is feasible even in a low-Q cavity. The scheme provides a flexible way for implementing nonreciprocal PB effects, and can be used as an unidirectional phonon source or memory in a chiral quantum network.

Appendix A: Derivation of the effective Hamiltonian

For a hybrid system consisting of a two-level atom and the mechanical mode both coupled to a cavity, the total Hamiltonian after linearization of the system can be written as

$$\begin{aligned} H_{2} & =H_{0}+H_{I},\cr\cr H_{0} & =(\Delta_c+\Delta_F)a^{\dagger}a+\Delta_a\sigma_{+}\sigma_{-}+\omega_mb^{\dagger}b,\cr\cr H_{I} & =J(a^{\dagger}\sigma_{-}+a\sigma_{+})+G(a^{\dagger}b+ab^{\dagger}). \end{aligned}$$

In the case of $|\delta _a|\gg J$ and $|\delta _b|\gg G$, with $\delta _a=\Delta _a-\Delta _c$, $\delta _b=\omega _m-\Delta _c$, the effective coupling between the two-level atom and the mechanical mode can be obtained by applying the Fröhlich-Nakajima transformation onto the Hamiltonian $H_{2}$ in Eq. (17).

We consider a unitary transformation $V$=exp($S$), where

$$S=\dfrac{J}{\delta_a+\Delta_{F}}(a^{\dagger}\sigma_{-}-a\sigma_{+})+\dfrac{G}{\delta_b+\Delta_{F}}(a^{\dagger}b-ab^{\dagger})$$
is anti-Hermitian and satisfies $H_{I}+[H_0,S]=0$. Up to the second order, the transformed Hamiltonian, $H=V^{\dagger }H_{2}V$, can be approximatively written as [60,61]
$$\begin{aligned} H & \approx H_{0}+\dfrac{1}{2}[H_{I},V]\cr\cr & =\left(\Delta_a+\dfrac{J^2}{\delta_a+\Delta_{F}}\right)\sigma_{+}\sigma_{-}+\left(\omega_m+\dfrac{G^2}{\delta_b+\Delta_{F}}\right)b^{\dagger}b\cr\cr & +\left(\dfrac{1}{2}\dfrac{JG}{\delta_a+\Delta_{F}}+\dfrac{1}{2}\dfrac{JG}{\delta_b+\Delta_{F}}\right)(b^{\dagger}\sigma_{-}+b\sigma_{+}) +\left(\Delta_c+\Delta_{F}-\dfrac{G^2}{\delta_b+\Delta_{F}}+\dfrac{J^2}{\delta_a+\Delta_{F}}\sigma{z}\right)a^{\dagger}a. \end{aligned}$$

In this expression the cavity has been decoupled from the atom-phonon system. Therefore, it is reasonable to assume that the cavity mode with large detuning always remains in the ground state, i.e., $\langle a^{\dagger }a\rangle \approx 0$. Under this approximation of $\langle a^{\dagger }a\rangle \approx 0$, the effective Hamiltonian in Eq. (19) can be reduced to Eq. (5) in the main text.

Appendix B: Effect of cavity driving on average photon number

In the above Eq. (19), since the cavity mode is decoupled from the system, it cannot influence the evolution of phonon-atom system. To demonstrate the validity of approximation $\langle a^{\dagger }a\rangle \approx 0$, we investigate the impact of optical driving field on the mean photon number.

By the standard linearization procedure with Hamiltonian in Eq. (2), the Heisenberg-Langevin equations for the average amplitudes of the system are given by

$$\begin{aligned} \dot{\bar{a}} & ={-}i(\Delta_{c}+\Delta_{F})\bar{a}-iJ\bar{\sigma}_{-}-ig\bar{a}(\bar{b}^{{\ast}}+\bar{b})-i\varepsilon-\kappa\bar{a}/2,\cr\cr \dot{\bar{b}} & ={-}i\omega_{m}\bar{b}-ig\vert{\bar{a}}\vert^2-\gamma \bar{b}/2,\cr\cr \dot{\bar{\sigma}}_{-} & ={-}i\Delta_a\bar{\sigma}_{-}-iJ\bar{\sigma}_{z}\bar{a}, \end{aligned}$$
where we have taken $\Delta _a=\omega _m$ so that $\delta _a=\delta _b=\delta$. Letting $\bar {\sigma }_{z}=-1$ (the atom in its ground state) for simplicity, and optomechanical coupling $g\ll \kappa$, the average amplitude of cavity mode can be calculated as
$$\bar{a}=\dfrac{i\varepsilon}{-i(\Delta_{c}+\Delta_{F})+iJ^2/\Delta_a-\kappa/2}.$$

The mean photon number $\langle a^{\dagger }a\rangle$ can be depicted in Fig. 9 using Eq. (21), where it tends to approach zero for large detuning. The approximation $\langle a^{\dagger }a\rangle \approx 0$ is valid. Nevertheless, it should be noted that in Fig. 9, the average photon number exhibits a monotonic increase as the drive amplitude of the cavity mode increases. Therefore, the drive amplitude of the cavity mode should not be too large in order to maintain the vacuum state of the cavity.

 figure: Fig. 9.

Fig. 9. The driving amplitude of the cavity mode $\varepsilon /\kappa$ as a function of cavity photon number $\langle a^{\dagger }a\rangle$ for different detuning frequency $\Delta _{c}$. The parameters are taken as $J/\kappa =5$, $\Delta _a/\kappa =\omega _{m}/\kappa =10$, $\Delta _{F}/\kappa =0$, $\gamma /\kappa =0.03$.

Download Full Size | PDF

Funding

National Natural Science Foundation of China (12264051).

Disclosures

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

Data availability

No data were generated or analyzed in the presented research.

References

1. A. Nunnenkamp, K. Børkje, and S. M. Girvin, “Single-photon optomechanics,” Phys. Rev. Lett. 107(6), 063602 (2011). [CrossRef]  

2. A. Kronwald, M. Ludwig, and F. Marquardt, “Full photon statistics of a light beam transmitted through an optomechanical system,” Phys. Rev. A 87(1), 013847 (2013). [CrossRef]  

3. P. Rabl, “Photon blockade effect in optomechanical systems,” Phys. Rev. Lett. 107(6), 063601 (2011). [CrossRef]  

4. D. Hu, S. Y. Huang, J. Q. Liao, L. Tian, and H. S. Goan, “Quantum coherence in ultrastrong optomechanics,” Phys. Rev. A 91(1), 013812 (2015). [CrossRef]  

5. H. Flayac and V. Savona, “Unconventional photon blockade,” Phys. Rev. A 96(5), 053810 (2017). [CrossRef]  

6. X. Y. Lü, Y. Wu, J. R. Johansson, H. Jing, J. Zhang, and F. Nori, “Squeezed optomechanics with phase-matched amplification and dissipation,” Phys. Rev. Lett. 114(9), 093602 (2015). [CrossRef]  

7. T. S. Yin, X. Y. Lü, L. L. Zheng, M. Wang, S. Li, and Y. Wu, “Nonlinear effects in modulated quantum optomechanics,” Phys. Rev. A 95(5), 053861 (2017). [CrossRef]  

8. J. Y. Yang, Z. Yang, C. S. Zhao, R. Peng, and L. Zhou, “Nonlinearity enhancement and photon blockade in hybrid optomechanical systems,” Opt. Express 29(22), 36167 (2021). [CrossRef]  

9. I. L. Aleiner, P. W. Brouwer, and L. I. Glazman, “Quantum effects in Coulomb blockade,” Phys. Rep. 358(5-6), 309–440 (2002). [CrossRef]  

10. S. Y. Li and A. D. Zhu, “Perfect single-magnon generator based on a hybrid cavity-magnonic system,” Ann. Phys. 534(5), 2100609 (2022). [CrossRef]  

11. Z. X. Liu, H. Xiong, and Y. Wu, “Magnon blockade in a hybrid ferromagnet-superconductor quantum system,” Phys. Rev. B 100(13), 134421 (2019). [CrossRef]  

12. X. Y. Yao, H. Ali, and P. B. Li, “Nonreciprocal phonon blockade in a spinning acoustic ring cavity coupled to a two-level system,” Phys. Rev. Appl. 17(5), 054004 (2022). [CrossRef]  

13. X. W. Xu, A. X. Chen, and Y. X. Liu, “Phonon blockade in a nanomechanical resonator resonantly coupled to a qubit,” Phys. Rev. A 94(6), 063853 (2016). [CrossRef]  

14. X. W. Xu, A. Miranowicz, H. R. Li, and F. Nori, “Method for observing robust and tunable phonon blockade in a nanomechanical resonator coupled to a charge qubit,” Phys. Rev. A 93(6), 063861 (2016). [CrossRef]  

15. M. Wang, T. S. Yin, Z. Y. Sun, H. G. Cheng, and L. L. Zheng, “Unconventional phonon blockade via atom-photon-phonon interaction in hybrid optomechanical systems,” Opt. Express 30(7), 10251 (2022). [CrossRef]  

16. J. Y. Yang, Z. Jin, J. S. Liu, H. F. Wang, and A. D. Zhu, “Unconventional phonon blockade in a Tavis–Cummings coupled optomechanical system,” Ann. Phys. 532(12), 2000299 (2020). [CrossRef]  

17. B. Sarma and A. K. Sarma, “Quantum-interference-assisted photon blockade in a cavity via parametric interactions,” Phys. Rev. A 96(5), 053827 (2017). [CrossRef]  

18. A. Miranowicz, J. Bajer, N. Lambert, Y. X. Liu, and F. Nori, “Tunable multiphonon blockade in coupled nanomechanical resonators,” Phys. Rev. A 93(1), 013808 (2016). [CrossRef]  

19. X. Wang, A. Miranowicz, H. R. Li, and F. Nori, “Multiple-output microwave single-photon source using superconducting circuits with longitudinal and transverse couplings,” Phys. Rev. A 94(5), 053858 (2016). [CrossRef]  

20. T. Ramos, V. Sudhir, K. Stannigel, P. Zoller, and T. J. Kippenberg, “Nonlinear quantum optomechanics via individual intrinsic two-level defects,” Phys. Rev. Lett. 110(19), 193602 (2013). [CrossRef]  

21. R. Fleury, D. Sounas, M. R. Haberman, and A. Alu, “Nonreciprocal acoustics,” Acoust. Today. 11, 14 (2015).

22. D. L. Sounas and A. Alu, “Non-reciprocal photonics based on time modulation,” Nat. Photonics 11(12), 774–783 (2017). [CrossRef]  

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

24. H. Nassar, B. Yousefzadeh, R. Fleury, M. Ruzzene, A. Alù, C. Daraio, A. N. Norris, G. Huang, and M. R. Haberman, “Nonreciprocity in acoustic and elastic materials,” Nat. Rev. Mater. 5(9), 667–685 (2020). [CrossRef]  

25. 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]  

26. H. Ramezani, P. K. Jha, Y. Wang, and X. Zhang, “Nonreciprocal localization of photons,” Phys. Rev. Lett. 120(4), 043901 (2018). [CrossRef]  

27. B. Peng, 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]  

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. K. Wang, Q. W Y. F. Yu, and Z. M. Zhang, “Nonreciprocal photon blockade in a two-mode cavity with a second-order nonlinearity,” Phys. Rev. A 100(5), 053832 (2019). [CrossRef]  

30. Q. T. Cao, H. M. Wang, C. H. Dong, H. Jing, R. S. Liu, X. Chen, L. Ge, Q. H. Gong, and Y. F. Xiao, “Experimental demonstration of spontaneous chirality in a nonlinear microresonator,” Phys. Rev. Lett. 118(3), 033901 (2017). [CrossRef]  

31. L. Fan, J. Wang, L. T. Varghese, H. Shen, B. Niu, Y. Xuan, A. M. Weiner, and M. H. Qi, “An all-silicon passive optical diode,” Science 335(6067), 447–450 (2012). [CrossRef]  

32. Y. Shi, Z. F. Yu, and S. H. Fan, “Limitations of nonlinear optical isolators due to dynamic reciprocity,” Nat. Photonics 9(6), 388–392 (2015). [CrossRef]  

33. N. Wang, Z. B. Yang, S. Y. Li, Y. L. Tong, and A. D. Zhu, “Nonreciprocal transmission and asymmetric entanglement induced by magnetostriction in a cavity magnomechanical system,” Eur. Phys. J. Plus 137(4), 422 (2022). [CrossRef]  

34. Z. B. Yang, J. S. Liu, A. D. Zhu, and R. C. Yang, “Nonreciprocal transmission and nonreciprocal entanglement in a spinning microwave magnomechanical system,” Ann. Phys. 532(9), 2000196 (2020). [CrossRef]  

35. S. Barzanjeh, M. Aquilina, and A. Xuereb, “Manipulation the flow of thermal noise in quantum devices,” Phys. Rev. Lett. 120(6), 060601 (2018). [CrossRef]  

36. 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,” Nat. London 558(7711), 569–572 (2018). [CrossRef]  

37. Y. F. Jiao, S. D. Zhang, Y. L. Zhang, A. Miranowicz, L. M. Kuang, and H. Jing, “Nonreciprocal optomechanical entanglement against backscattering losses,” Phys. Rev. Lett. 125(14), 143605 (2020). [CrossRef]  

38. Y. F. Jiao, J. X. Liu, Y. Li, L. M. Kuang, and H. Jing, “Nonreciprocal enhancement of remote entanglement between nonidentical mechanical oscillators,” Phys. Rev. Appl. 18(6), 064008 (2022). [CrossRef]  

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

40. B. Li, X. W. Xu, L. Zhang, L. M. Kuang, and H. Jing, “Nonreciprocal optical solitons in a spinning Kerr resonator,” Phys. Rev. A 103(5), 053522 (2021). [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 (2019). [CrossRef]  

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

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

44. H. L. Zhang, R. Huang, S. D. Zhang, Y. Li, C. W. Qiu, F. Nori, and H. Jing, “Breaking anti-PT symmetry by spinning a resonator,” Nano Lett. 20(10), 7594–7599 (2020). [CrossRef]  

45. R. Huang, A. Miranowicz, J. Q. Liao, F. Nori, and H. Jing, “Nonreciprocal photon blockade,” Phys. Rev. Lett. 121(15), 153601 (2018). [CrossRef]  

46. X. W. Xu, Y. J. Zhao, H. Wang, H. Jing, and A. X. Chen, “Nonreciprocal photon blockade via quadratic optomechanical coupling,” Photonics Res. 8(2), 143 (2020). [CrossRef]  

47. Y. M. Wang, W. Xiong, Z. Y. Xu, G. Q. Zhang, and J. Q. You, “Dissipation-induced nonreciprocal magnon blockade in a magnon-based hybrid system,” Sci. China Phys. Mech. Astron. 65(6), 260314 (2022). [CrossRef]  

48. W. S. Xue, H. Z. Shen, and X. X. Yi, “Nonreciprocal conventional photon blockade in driven dissipative atom-cavity,” Opt. Lett. 45(16), 4424 (2020). [CrossRef]  

49. B. J. Li, R. Huang, X. W. Xu, A. Miranowicz, and H. Jing, “Nonreciprocal unconventional photon blockade in a spinning optomechanical system,” Photonics Res. 7(6), 630 (2019). [CrossRef]  

50. M. Ludwig, A. H. Safavi-Naeini O. Painter, and F. Marquardt, “Enhanced quantum nonlinearities in a two-mode optomechanical system,” Phys. Rev. Lett. 109(6), 063601 (2012). [CrossRef]  

51. K. Stannigel, P. Komar, P. Zoller, and P. Rabl, “Optomechanical quantum information processing with photons and phonons,” Phys. Rev. Lett. 109(1), 013603 (2012). [CrossRef]  

52. N. Lörch and K. Hammerer, “Sub-Poissonian phonon lasing in three-mode optomechanics,” Phys. Rev. A 91(6), 061803 (2015). [CrossRef]  

53. I. S. Grudinin, H. Lee, O. Painter, and K. J. Vahala, “Phonon laser action in a tunable two-level system,” Phys. Rev. Lett. 104(8), 083901 (2010). [CrossRef]  

54. H. Wang, Z. X. Wang, J. Zhang, S. K. Özdemir, and Y. X. Liu, “Phonon amplification in two coupled cavities containing one mechanical resonator,” Phys. Rev. A 90(5), 053814 (2014). [CrossRef]  

55. H. J. Chen, X. W. Fang, C. Z. Chen, and Y. Li, “Coherent optical propagation and ultrahigh resolution mass sensor based on photonic molecules optomechanics,” Opt. Commun. 382, 73–79 (2017). [CrossRef]  

56. J. M. Silver, L. D. Bino, S. Y. Zhang, M. R. Vanner, and A. Ø Svela, “Coherent suppression of backscattering in optical microresonators,” Light: Sci. Appl. 9(1), 204 (2020). [CrossRef]  

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

58. J. X. Liu, Y. F. Jiao, Y. Li, X. W. Xu, Q. Y. He, and H. Jing, “Phase-controlled asymmetric optomechanical entanglement against optical backscattering,” Sci. China Phys. Mech. Astron 66(3), 230312 (2023). [CrossRef]  

59. G. B. Malykin, “The Sagnac effect: correct and incorrect explanations,” Phys. Usp. 43(12), 1229–1252 (2000). [CrossRef]  

60. H. Fröhlich, “Theory of the superconducting state. I. The ground state at the absolute zero of temperature,” Phys. Rev. 79(5), 845–856 (1950). [CrossRef]  

61. S. Nakajima, “Perturbation theory in statistical mechanics,” Adv. Phys. 4(16), 363–380 (1955). [CrossRef]  

62. J. Johansson, P. D. Nation, and F. Nori, “QuTiP: An open-source Python framework for the dynamics of open quantum systems,” Comput. Phys. Commun. 183(8), 1760–1772 (2012). [CrossRef]  

63. J. Johansson, P. D. Nation, and F. Nori, “QuTiP2: A Python framework for the dynamics of open quantum system,” Comput. Phys. Commun. 184(4), 1234–1240 (2013). [CrossRef]  

64. V. Huet, A. Rasoloniaina, P. Rochard, M. Mortier, A. Yacomotti, and Y. Dumeige, “Millisecond photon lifetime in a slow-light microcavity,” Phys. Rev. Lett. 116(13), 133902 (2016). [CrossRef]  

65. R. Reimann, M. Doderer, E. Hebestreit, M. Frimmer, D. Windey, F. Tebbenjohanns, and L. Novotny, “GHz rotation of an optically trapped nanoparticle in vacuum,” Comput. Phys. Commun. 121(3), 033602 (2018). [CrossRef]  

66. J. Ahn, Z. Xu, J. Bang, Y. H. Deng, T. M. Hoang, Q. Han, R. M. Ma, and T. Li, “Optically levitated nanodumbbell torsion balance and GHz nanomechanical rotor,” Phys. Rev. Lett. 121(3), 033603 (2018). [CrossRef]  

67. P. Komar, S. D. Bennett, K. Stannigel, S. J. M. Habraken, P. Rabl, P. Zoller, and M. D. Lukin, “Single-photon nonlinearities in two-mode optomechanics,” Phys. Rev. A 87(1), 013839 (2013). [CrossRef]  

68. H. Wang, X. Gu, Y. x. Liu, A. Miranowicz, and F. Nori, “Tunable photon blockade in a hybrid system consisting of an optomechanical device coupled to a two-level system,” Phys. Rev. A 92(3), 033806 (2015). [CrossRef]  

69. P. Z. G. Fonseca, E. B. Aranas, J. Millen, and P. F. Barker, “Nonlinear dynamics and strong cavity cooling of levitated nanoparticles,” Phys. Rev. Lett. 117(17), 173602 (2016). [CrossRef]  

70. C. Hamsen, K. N. Tolazzi, T. Wilk, and G. Rempe, “Two-photon blockade in an atom-driven cavity QED system,” Phys. Rev. Lett. 118(13), 133604 (2017). [CrossRef]  

71. C. Xiong, X. K. Sun, and H. X. Tang, “Integrated high frequency aluminum nitride optomechanical resonators,” Appl. Phys. Lett. 100(17), 171111 (2012). [CrossRef]  

Data availability

No data were generated or analyzed in the presented research.

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 (9)

Fig. 1.
Fig. 1. A spinning optomechanical resonator coupled to a two-level atom. Spinning the resonator results in different Fizeau drag $\Delta _F$ for engineering the clockwise or counterclockwise optical mode undergoing different refractive indices with an angular velocity $\Omega$. (a) denotes driving the resonator from left side ($\Delta _F>0$), and (b) from right side ($\Delta _F<0$).
Fig. 2.
Fig. 2. The energy spectrum of the phonon-atom subsystem in the few-phonon subspace when the angular velocity $\Omega =0$. $|m,g(e)\rangle$ represents the state with $m$ phonons and the atom in ground (excited) state. The green and purple arrows represent the frequency of driving laser.
Fig. 3.
Fig. 3. Logarithmic second-order correlation function $\lg {g^{(2)}(0)}$ in (a) and the mean phonon number $N_{b}$ in (c) as functions of $J/\kappa$ and $\Delta /\kappa$ for $G/\kappa =10$. Logarithmic second-order correlation function $\lg {g^{(2)}(0)}$ in (b) and mean phonon number $N_{b}$ in (d) as functions of $G/\kappa$ and $\Delta /\kappa$ for $J/\kappa =10$. The white dotted line denotes the optimal conditions of the conventional PB in Eq. (10). The red contour denotes the case of second-order correlation function $g^{(2)}(0)=1$. The other parameters are taken as $\Omega =0$, $E/\kappa =0.01$, $\delta /\kappa =50$, $n_{th}=0$, $\gamma /\kappa =0.03$.
Fig. 4.
Fig. 4. (a) Second-order correlation function $g^{(2)}{(0)}$ versus the detuning $\Delta /\kappa$ for different driving directions of the cavity. The nonreciprocal 1PB occurs in solved detuning region at $\Delta /\kappa =3.70711$ and $\Delta _{F}/\kappa =\pm 3.95043$. (b) The energy levels when driving the system from left and right side is nonequidistant for $\Delta /\kappa =3.70711$. Other parameters are the same as those in Fig. 3.
Fig. 5.
Fig. 5. The energy level spectrum of nonreciprocal 2PB. With enhanced driving strength $E/\kappa =0.2$ and the same Fizeau drag $\Delta _{F}/\kappa =\pm 3.95043$, 2PB occurs when the resonator is driven from left side ($\Delta _{F}>0$), while three-phonon induced PIT happens when the resonator is driven from right side ($\Delta _{F}<0$).
Fig. 6.
Fig. 6. (a) Second-order correlation function $g_{L,R}^{(2)}(0)$ (solid curves) and three-order correlation function $g_{L,R}^{(3)}(0)$ (dashed curves) versus the frequency detuning for different driving directions, and the nonreciprocal 2PB can be observed clearly by the enlarged figure in (b). Relative deviation of the phonon population from the Poisson distribution is shown in (c) when the nonreciprocal 2PB occurs and in (d) when the PIT occurs.
Fig. 7.
Fig. 7. The numerical results of second-order correlation function $g^{(2)}(0)$ simulated by the linearized Hamiltonian $H_l$ and the effective Hamiltonian $H_{\rm {eff}}$ coincide well with each other. The parameters are taken as $\tilde {\Delta }_{c}/\kappa =50$, $J/\kappa =G/\kappa =10$, $E/\kappa =0.01$, $\gamma /\kappa$=0.03.
Fig. 8.
Fig. 8. The second-order correlation function $g^{(2)}(0)$ as a function of the detuning $\Delta /\kappa$ for different thermal phonon numbers $n_{th}$. Other parameters are the same as those in Fig. 3.
Fig. 9.
Fig. 9. The driving amplitude of the cavity mode $\varepsilon /\kappa$ as a function of cavity photon number $\langle a^{\dagger }a\rangle$ for different detuning frequency $\Delta _{c}$. The parameters are taken as $J/\kappa =5$, $\Delta _a/\kappa =\omega _{m}/\kappa =10$, $\Delta _{F}/\kappa =0$, $\gamma /\kappa =0.03$.

Equations (21)

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

Δ F = ± n r Ω ω c c ( 1 1 n 2 1 n d n d λ ) ,
H = ( Δ c + Δ F ) a a + Δ a σ + σ + ω m b b + J ( a σ + a σ + ) + g a a ( b + b ) + ε ( a + a ) ,
H 1 = ( Δ c + Δ F ) a a + Δ a σ + σ + ω m b b + J ( a σ + a σ + ) + g a ¯ ( a + a ) ( b + b ) ,
H 2 = ( Δ c + Δ F ) a a + Δ a σ + σ + ω m b b + J ( a σ + a σ + ) + G ( a b + a b ) ,
H 3 = ( Δ a + J 2 δ + Δ F ) σ + σ + ( ω m + G 2 δ + Δ F ) b b + ( J G δ + Δ F ) ( b σ + b σ + ) ,
H e f f = ( Δ + J 2 δ + Δ F ) σ + σ + ( Δ + G 2 δ + Δ F ) b b + ( J G δ + Δ F ) ( b σ + b σ + ) + E ( b + b ) ,
H = ( m ( Δ + G 2 δ + Δ F )             J G δ + Δ F m J G δ + Δ F m             m ( Δ + J 2 δ + Δ F ) ) .
E m , ± = m [ J 2 + G 2 2 Δ ( δ + Δ F ) ] ± m K 2 ( δ + Δ F ) ,
| m , ± = m ( G 2 J 2 ) ± K 2 J G | m , g + | m 1 , e ,
J 2 = G 2 + Δ ( δ + Δ F ) .
ρ ˙ = i [ H e f f , ρ ] + γ a 2 ( 2 σ ρ σ + σ + σ ρ ρ σ + σ ) + γ m 2 ( n t h + 1 ) ( 2 b ρ b b b ρ ρ b b ) + γ m 2 n t h ( 2 b ρ b b b ρ ρ b b ) ,
g ( 2 ) ( 0 ) = T r ( ρ b b b b ) [ T r ( ρ b b ) ] 2 = b b b b b b 2 .
g ( 2 ) ( 0 ) > 1 & g ( 3 ) ( 0 ) < 1.
[ P ( m ) P ( m ) ] / P ( m ) ,
H l = Δ ~ c a a + Δ σ + σ + Δ b b + J ( a σ + a σ + ) + G ( a b + a b ) + E ( b + b ) ,
ρ ˙ = i [ H l , ρ ] + κ 2 ( 2 a ρ a a a ρ ρ a a ) + γ 2 ( 2 σ ρ σ + σ + σ ρ ρ σ + σ ) + γ 2 ( n t h + 1 ) ( 2 b ρ b b b ρ ρ b b ) + γ 2 n t h ( 2 b ρ b b b ρ ρ b b ) ,
H 2 = H 0 + H I , H 0 = ( Δ c + Δ F ) a a + Δ a σ + σ + ω m b b , H I = J ( a σ + a σ + ) + G ( a b + a b ) .
S = J δ a + Δ F ( a σ a σ + ) + G δ b + Δ F ( a b a b )
H H 0 + 1 2 [ H I , V ] = ( Δ a + J 2 δ a + Δ F ) σ + σ + ( ω m + G 2 δ b + Δ F ) b b + ( 1 2 J G δ a + Δ F + 1 2 J G δ b + Δ F ) ( b σ + b σ + ) + ( Δ c + Δ F G 2 δ b + Δ F + J 2 δ a + Δ F σ z ) a a .
a ¯ ˙ = i ( Δ c + Δ F ) a ¯ i J σ ¯ i g a ¯ ( b ¯ + b ¯ ) i ε κ a ¯ / 2 , b ¯ ˙ = i ω m b ¯ i g | a ¯ | 2 γ b ¯ / 2 , σ ¯ ˙ = i Δ a σ ¯ i J σ ¯ z a ¯ ,
a ¯ = i ε i ( Δ c + Δ F ) + i J 2 / Δ a κ / 2 .
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.