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

Nonreciprocal $\mathcal {PT}$-symmetric magnon laser in spinning cavity optomagnonics

Open Access Open Access

Abstract

We propose a scheme to achieve nonreciprocal parity-time ($\mathcal {PT}$)-symmetric magnon laser in a $\mathcal {PT}$-symmetric cavity optomagnonical system. The system consists of active and passive optical spinning resonators. We demonstrate that the Fizeau light-dragging effect induced by the spinning of a resonator results in significant variations in magnon gain and stimulated emitted magnon numbers for different driving directions. We find that utilizing the Fizeau light-dragging effect allows the system to operate at ultra-low thresholds even without reaching gain-loss balance. A one-way magnon laser can also be realized across a range of parameters. High tunability of the magnon laser is achieved by changing the spinning speed of the resonators and driving direction. Our work provides a new way to explore various nonreciprocal effects in non-Hermitian magnonic systems, which may be applied to manipulate photons and magnons in multi-body non-Hermitian coupled systems.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent years, cavity optomagnonical system has garnered significant attention in quantum physics due to the strong or even ultrastrong couplings of collective spin excitations in a millimeter-scale yttrium iron garnet (YIG) crystal and the cavity photons [17]. The utilization of YIG sphere is universal in hybrid systems based on magnonics, primarily owing to their remarkable advantages of high spin density and low damping rate [8]. Strong and even ultrastrong coupling between the magnon mode in the ferromagnet YIG sphere and the cavity photons has been proposed in Refs. [913]. The cavity optomagnonical system has become a promising platform for the implementation of various interesting phenomena, such as magnon dark modes memory [14], chaos [15], exceptional point [1618], dissipative magnon-photon coupling [1921] and stationary one-way quantum steering [22,23].

The concept of parity-time ($\mathcal {PT}$)-symmetry has recently attracted a lot of attention. It is well-understood that when system parameters in the Hamiltonian are appropriately adjusted, a transition occurs from an unbroken $\mathcal {PT}$-symmetric phase (with a real eigenvalue spectrum) to a spontaneously broken $\mathcal {PT}$-symmetric phase (characterized by a complex eigenvalue spectrum) [24]. This shift usually occurs around the exceptional points (EPs) [25,26], i.e., $\mathcal {PT}$-symmetric phase transition point [2729]. This phenomenon attracted remarkable interests and rapidly emerged as a forefront topic in the area of non-Hermitian physics. In the field of cavity optomagnonics, these advancements have resulted in a range of important applications, including the realization of magnon-induced chaos in a $\mathcal {PT}$-symmetric cavity optomagnonical system [30]. Moreover, the observations of EPs in non-Hermitian systems [16,3133], as well as the controlled EPs and bound states in anti-$\mathcal {PT}$-symmetric cavity optomagnonical systems have been already investigated [34]. Importantly, a $\mathcal {PT}$-symmetric magnon laser in a cavity optomagnonical system has been proposed in Ref. [35] recently. This development signifies that $\mathcal {PT}$-symmetric systems have indeed emerged as a promising platform for investigating magnon laser effects.

Nonreciprocal devices have the ability to allow incident light to exhibit different properties in different propagation directions, which play a crucial role in a variety of applications, for instance signal processing and invisibility sensing [36]. Several physical mechanisms such as optomechanics [37,38], non-Hermitian optics [39,40], and spinning resonators [4144] have become great platforms for studying optical nonreciprocity. Furthermore, the nonreciprocal phonon laser by coupling optomechanical resonators with spinning optical resonators has been proposed in Ref. [45]. A scheme for generating nonreciprocal phonon lasers in a single spinning mechanical resonator is realized [46], extending nonreciprocal phonon laser to cavity magnomechanical system.

So far, based on the Brillouin scattering process in YIG sphere, magnon laser have been theoretically proposed as a new type of laser similar to optical laser and phonon laser, and the experimental feasibility under current experimental conditions has been evaluated theoretically [47]. The magnon laser has quickly become an important platform for research the intersections of quantum optics and magnon spintronics. Note that nonreciprocal magnon laser was recently achieved by coupling to a spinning resonator [48]. In addition, the nonreciprocal magnon laser realized by introducing parametric amplification further improved the nonreciprocity [49,50]. Recently, a $\mathcal {PT}$-symmetric magnon laser has been achieved in a cavity optomagnonical system through the introduction of gain [35]. This achievement significantly reduces the threshold power of the magnon laser and extends its applicability to non-Hermitian domains. Notably, the exploration of nonreciprocal magnon lasers within $\mathcal {PT}$-symmetric systems also remains an underexplored area, despite the substantial of efforts devoted to magnon lasers.

Inspired by the implications of these studies, in this paper, we investigate the nonreciprocal $\mathcal {PT}$-symmetric magnon laser consists of active and passive optical whispering gallery mode (WGM) spinning resonators and a YIG sphere, where the magnon mode in YIG sphere couples with the cavity mode $a_2$ in passive resonator. We demonstrate that when driving the system from one direction enhances the magnon laser, while driving it from the opposite direction suppresses the magnon laser. The Fizeau light-dragging effect, generated by the resonator spinning, adjusts the steady-state photon population, magnon gain, and threshold power in the two resonators. Therefore, by tuning the spinning speed of the resonators and driving direction, a highly tunable nonreciprocal $\mathcal {PT}$-symmetric magnon laser can be achieved. Finally, by increasing the active cavity gain coefficient, we obtain a nonreciprocal magnon laser with ultra-low threshold power, and one-way magnon laser can be realized across a range of input power. Compared with earlier studies on $\mathcal {PT}$-symmetric magnon lasers [35], our work achieves nonreciprocity in the device by introducing spinning cavities of gain and loss, which allows our work to realize a unidirectional magnon laser in $\mathcal {PT}$-symmetric systems. Besides, in contrast to previous nonreciprocal magnon laser scheme [49], our work significantly increases the maximum stimulated emission magnons number of the system and decreases the threshold power. Our scheme provides a new way to explore various nonreciprocal effects in non-Hermitian cavity magnonics systems and holds potential applications in various spintronics devices.

The paper is organized as follows: In Sec. 2, we introduce the physical model and the Hamiltonian of the $\mathcal {PT}$-symmetric cavity optomagnonical system, and describe the system dynamics using the Heisenberg-Langevin equations. We analyze the position of the EP, determine the optimal parameters of the system, and obtain the magnon gain using an adiabatic elimination method. In Sec. 3, we extensively discuss the nonreciprocity of the system, including the variations of the steady-state photon populations, the magnon gain, the stimulated emitted magnon number, and the threshold power for different driving directions. Finally, a conclusion is given in Sec. 4.

2. System and Hamiltonian

The $\mathcal {PT}$-symmetric cavity optomagnonical system, which consists of a active WGM spinning resonator with doped gain medium, a passive WGM spinning resonator and a ferrimagnet YIG sphere as shown in Fig. 1. The YIG sphere is uniformly magnetized along the $\mathit {z}$ direction by an external static magnetic field $\textbf {B}_\mathit {z}$ = $B\textbf {e}_\mathit {z}$. Two optical WGMs in spinning resonators sharing the same resonant frequency $\omega _\textrm {c}$, refractive index, radius. Both resonators are equipped with spinning bases that rotate in opposite directions. Therefore, the clockwise and counter-clockwise optical WGMs in the spinning resonators have different refractive indices, which causes them to experience a Fizeau shift, the key factor leading to nonreciprocal behavior. The Fizeau shift $\Delta _{\mathrm {F}}$ can be written as [51]

$$\Delta_{\mathrm{F}}={\pm} \Omega \frac{n r \omega_\textrm{c}}{c}\left(1-\frac{1}{n^2}-\frac{\lambda}{n} \frac{d n}{d \lambda}\right),$$
where $\Omega$ represents the angular velocity of the two spinning resonators, $\textit {n}$ and $\textit {r}$ denote their refractive index and radius, respectively. $\textit {c}$ represents the speed of light in vacuum, $\lambda$ denotes the wavelength of light. The dispersion term $(\lambda /{n}) ({d n}/{d \lambda })$ represents the Lorentz correction of the Fresnel-Fizeau drag coefficient, characterizing the relativistic origin of the Sagnac effect. This term is usually small and can be neglected. Maintaining the rotational direction of the spinning resonators constant, $\Delta _{\mathrm {F}} > 0$ ($\Delta _{\mathrm {F}} < 0$) indicates that the active resonator is excited from the left (right) side, where the pumping light input direction is the same (opposite) to the resonators rotational direction.

 figure: Fig. 1.

Fig. 1. Schematic diagram of a nonreciprocal $\mathcal {PT}$-symmetric cavity optomagnonical system. The ferrimagnet YIG sphere supports optical WGM $a_{2}$ and magnon mode $m$ and optical WGM $a_{2}$ couples to active optical WGM $a_{1}$. The input light evanescent couples to the active optical WGM via a tapered fiber. Driving the system from the left ($\Delta _{\mathrm {F}}>0$) enhances the magnon laser, while driving from the right ($\Delta _{\mathrm {F}}<0$) suppresses or even blocks magnon laser emission.

Download Full Size | PDF

The total Hamiltonian of the system in the frame rotating at the cavity driving frequency $\omega _\textrm {d}$ can be written as

$$\begin{aligned} H= & ~H_{\textrm{0}}+H_{\textrm{int}}+H_{\textrm{dr}},\\ H_0= & ~\hbar (-\Delta-\Delta_{\mathrm{F}})(a_1^{\dagger} a_1+a_2^{\dagger} a_2)+\hbar\omega_\textrm{m}m^{\dagger}m,\\ H_{\textrm{int}}= & ~{\hbar}J\left(a_1^{\dagger}a_2+a_2^{\dagger}a_1\right)+\hbar\textit{G}\sqrt{\textit{S}/2}a_2^{\dagger}a_2(m+m^{\dagger}),\\ H_{\textrm{dr}}= & ~i \hbar \sqrt{\kappa / 2} \varepsilon_\textrm{d}\left(a_1^{\dagger}-a_1\right), \end{aligned}$$
where $H_0$ is the free Hamiltonian for the active optical WGM $a_1$, the passive optical WGM $a_2$, and the magnon mode $m$. The active optical WGM $a_1$ is fabricated from silica doped with $\textrm {Er}^{3+}$ ions, and the optical gain can be acquired by optically pumping the $\textrm {Er}^{3+}$ ions [40,52]. Both the active optical WGM $a_1$ and passive optical WGM $a_2$ sharing the same resonant frequency $\omega _\textrm {c}$. $\Delta =\omega _\textrm {d}-\omega _\textrm {c}$ is the frequency detuning between the driving light and the two optical WGMs. $\omega _\textrm {m}$ represents the frequency of magnon mode. $H_{\textrm {int}}$ describes the interaction Hamiltonian, the first term denotes the coupling of optical WGM $a_1$ to optical WGM $a_2$ through the optical tunneling effect with a coupling strength $\textit {J}$. The second term denotes the coupling between optical WGM $a_2$ and magnon mode $m$, characterized by a coupling strength denoted as $G=\frac {1}{S} \frac {c \vartheta _F}{4 \sqrt {\varepsilon }}$ [53], also referred to as the parametric optomagnonic coupling, where $S$ represents the total spin number of the YIG sphere, $\vartheta _F$ is the Faraday rotation per unit length, and $\varepsilon$ is the relative permittivity. The pumping light applied to the active optical WGM $a_1$ is considered as the $H_{\textrm {dr}}$ with the amplitude $\varepsilon _\textrm {d}=\sqrt {P_{\rm in}/(\hbar \omega _\textrm {d})}$, in which $\omega _\textrm {d}$ is the driving frequency and $P_{\rm in}$ represents the input power.

The Heisenberg-Langevin equations of motion of the system can be written as

$$\begin{aligned} \dot{a}_1= & ~(i\Delta+i\Delta_{\mathrm{F}}+\kappa / 2 )a_1-i J a_2+\sqrt{\kappa / 2} \varepsilon_\textrm{d},\\ \dot{a}_2= & ~(i\Delta+i\Delta_{\mathrm{F}}-\gamma / 2) a_2-i J a_1-i G \sqrt{S / 2} a_2\left(m^{\dagger}+m\right),\\ \dot{m}= & ~-i \omega_\textrm{m} m-\gamma_\textrm{m} / 2 m-i G \sqrt{S / 2} a_2^{\dagger} a_2, \end{aligned}$$
$\kappa$, $\gamma$, and $\gamma _\textrm {m}$ represent the gain of optical WGM $a_1$, the decay rate of optical WGM $a_2$, and the decay rate of the magnon mode, respectively. In the semiclassical approximation, taking into account the average response of the system, all operators are reduced to their expectation values thus the quantum correlations and noise terms can be safely ignored. The steady-state mean values of the system can be obtained as follows:
$$\begin{aligned} a_{1 s}= & ~\frac{i J a_{2s}-\sqrt{\kappa / 2} \varepsilon_\textrm{d}} {i \tilde{\Delta}+\kappa/2},\\ a_{2 s}= & ~\frac{4 J \sqrt{\kappa / 2} \varepsilon_\textrm{d}} {i({-}4 \tilde{\Delta}^2+J^2- \kappa \gamma)+2 (\gamma+2i \phi_0- \kappa)\tilde{\Delta}+2 \kappa \phi_0},\\ m_s= & ~\frac{-i G \sqrt{S / 2}\left|a_{2 s}\right|^2} {\left(i \omega_\textrm{m}+\gamma_\textrm{m} / 2\right)}, \end{aligned}$$
where $\phi _0=G \sqrt {S / 2}(m_s^*+m_s)$ and $\tilde {\Delta }=\Delta +\Delta _{\mathrm {F}}$. Before further investigation, it is necessary to illustrate the influence of EP on the $\mathcal {PT}$-symmetric system and determine the location of EP. Accordingly we investigate the $\mathcal {PT}$-symmetric dual-cavity system consists of a active optical WGM $a_1$ and a passive optical WGM $a_2$ in two spinning resonators, the non-Hermitian Hamiltonian of the system can be obtained by phenomenologically introducing imaginary gain and loss terms. Because of the newly introduced terms do not affect any of the previous calculations, the final reduced non-Hermitian Hamiltonian is
$$H_{\textrm{non}}=\hbar (-\tilde{\Delta}+i \kappa/2)a_1^{\dagger} a_1+\hbar (-\tilde{\Delta}-i \gamma/2)a_2^{\dagger}a_2+{\hbar}J\left(a_1^{\dagger}a_2+a_2^{\dagger}a_1\right),$$
it can be straightforwardly diagonalized as $\hbar (\omega _++i \gamma _+)\Psi _{+}^{\dagger } \Psi _{+}+\hbar (\omega _-+i \gamma _-)\Psi _{-}^{\dagger } \Psi _{-}$ via a Bogoliubov transformation, where $\omega _{\pm }=-\tilde {\Delta }\pm \sqrt {J^2-(\kappa +\gamma )^2/16}$ and $\gamma _{\pm }=(\kappa -\gamma )/4$, which is achieved by solving the eigenvalues of the Hamiltonian and corresponds to the eigenfrequencies and damping rates of the two eigenmodes $\Psi _{\pm }=(a_1\pm a_2)/\sqrt {2}$, i.e., optical supermodes. In order to reveal the positions of the EP and the mode splitting as well as linewidth of the supermode, we plot the eigenvalue spectrum of the non-Hermitian Hamiltonian mentioned above in Fig. 2. As shown Figs. 2(a) and 2(b), we can observe that when $J<(\kappa +\gamma )/4$, the eigenfrequencies of supermodes become degenerate while decay rates split. This implies that the system is in the $\mathcal {PT}$-symmetry broken phase (blue region). On the other hand, when $J>(\kappa +\gamma )/4$ the eigenfrequencies of supermodes begin to splitting, while the decay rates become degenerate, signifying the system enters the $\mathcal {PT}$-symmetric phase. Therefore, we can pinpoint $J=(\kappa +\gamma )/4$ as the position of the EP for the system. The eigenfrequencies splitting width $\Delta _{\omega }$ of the two supermodes $\Psi _{\pm }$ is $2\sqrt {J^2-(\kappa +\gamma )^2/16}$. Similar to the generation of optical lasers in a two-level system, the optimal matching condition for lasers requires the eigenfrequencies splitting width between two supermodes resonant with magnon frequency, i.e., $\Delta \omega =\omega _{\textrm {m}}$. Therefore, we can find the optimal intercavity coupling strength $J=\sqrt {(\kappa +\gamma )^2/16+\omega _{\textrm {m}}^2/4}$.

 figure: Fig. 2.

Fig. 2. (a) Eigenfrequencies splitting of supermodes $(\omega _{\pm }-\tilde {\Delta })/\gamma$ and (b) system dissipation of the eigenmodes $\gamma _{\pm }/\gamma$ vary with the optical tunneling rate $J/\gamma$ under the condition of $\kappa /\gamma$ = 1.

Download Full Size | PDF

By introducing the supermode operators, and applying the rotating-wave approximation, the Hamiltonian of the system can be rewritten as

$$\begin{aligned} \mathcal{H}= & \hbar \omega_{+} \Psi_{+}^{\dagger} \Psi_{+}+\hbar \omega_{-} \Psi_{-}^{\dagger} \Psi_{-}+\hbar \omega_\textrm{m} m^{\dagger} m \\ & -\frac{\hbar G \sqrt{S/2}}{2}\left(\Psi_{+}^{\dagger} \Psi_{-} m+m^{\dagger} \Psi_{-}^{\dagger} \Psi_{+}\right), \end{aligned}$$
the first and second terms describe the Hamiltonian of the supermodes $\Psi _{+}$ and $\Psi _{-}$ with the frequencies $\omega _{+}$ and $\omega _{-}$, the third term describes the free energy of the magnon mode, and the final term denotes the absorption and emission of magnon between two nondegenerate supermodes $\Psi _{\pm }$.

According the Hamiltonian as aforementioned, the dynamical equations for the supermodes $\Psi _{\pm }$ and magnon mode $m$ can be obtained from Eq. (6):

$$\begin{aligned} \dot{\Psi}_{+}= & ~(i\omega_{+}+\kappa/2)\Psi_{+}+iG\frac{\sqrt{S / 2}}{2}\Psi_{-}m,\\ \dot{\Psi}_{-}= & ~(i\omega_{-}-\gamma/2)\Psi_{-}+iG\frac{\sqrt{S / 2}}{2}\Psi_{+}m^{\dagger},\\ \dot{m}= & ~(i\omega_{m}-\gamma_{m}/2)m+iG\frac{\sqrt{S / 2}}{2}\Psi_{-}^{\dagger}\Psi_{+}. \end{aligned}$$

We can define population-inversion operator $\Delta _N=\Psi _{+}^{\dagger }\Psi _{+}-\Psi _{-}^{\dagger } \Psi _{-}$ and the ladder operator $p=\Psi _{-}^{\dagger }\Psi _{+}$ of the optical supermodes, the above dynamical equations can be rewritten as

$$\begin{aligned} & \dot{m}=\left({-}i \omega_m-\gamma_m / 2\right) m+i G \sqrt{S / 2} / 2 p, \\ & \dot{p}=\left({-}i \Delta \omega-\gamma^{\prime} / 2\right) p-i G \sqrt{S / 2} \Delta_N / 2 m , \end{aligned}$$
where $\gamma ^{\prime }=\kappa -\gamma$, transferring the variables $m$ and $p$ to a rotating frame through the execution of a unitary transformation $U(t)=$exp$(i\omega _mm^{\dagger }mt+i\Delta _{\omega }p^{\dagger }pt)$, we can obtain
$$\tilde{m}=U^{\dagger} m U=m e^{i \omega_m t}, \quad \tilde{p}=U^{\dagger} p U=p e^{i \Delta_\omega t}.$$

The equations of motion of the system then become

$$\dot{\tilde{m}}={-}\gamma_m / 2 \tilde{m}+i G \sqrt{S / 2} / 2 \tilde{\Phi} e^{i\left(\omega_m-\Delta \omega\right) t},$$
$$\dot{\tilde{\Phi}}={-}\gamma^{\prime} / 2 \tilde{\Phi}-i G \sqrt{S / 2} \Delta N / 2 \tilde{m} e^{i\left(\Delta \omega-\omega_m\right) t}.$$

Due to the optical decay is more rapid than the magnon decay, i.e., $\gamma, \gamma ^{\prime }\gg \gamma _{m}$, $\tilde {m}$ can be viewed as a constant compared to $\tilde {p}$ and since $\gamma ^{\prime }\gg G \sqrt {S}$, the time-dependent term in $\tilde {p}$ can be safely ignored [27,35]. Therefore, we can solve Eq. (10a) analytically with a first-order ordinary differential equation, obtaining the first-order approximate solution as:

$$\tilde{p}=\frac{-i G \sqrt{S / 2} \Delta_N / 2 e^{i\left(\Delta_\omega-\omega_m\right) t}}{i\left(\Delta_\omega-\omega_m\right)+\frac{\gamma^{\prime}}{2}} \tilde{m}.$$

By substituting $\tilde {p}$ into Eq. (10b) , the amplitude of magnon $\tilde {p}$ can be obtained as

$$\tilde{m}=\exp \left[\left(-\gamma_m / 2+\mathcal{G}-i \Omega_m\right) t\right].$$

From Eq. (12), the effective magnon gain $\mathcal {G}$ is given by

$$\mathcal{G}=\frac{G^2 S \Delta_N \gamma^{\prime} / 16}{\left(\Delta_{\omega}-\omega_\textrm{m}\right)^2+\left(\frac{\gamma^{\prime}}{2}\right)^2},$$

The magnon gain $\mathcal {G}$ represents the extent to which magnons are amplified through their interaction with stimulated emission when they transition from a higher energy level ($\omega _{+}$) to a lower energy level ($\omega _{-}$). Furthermore, because $\Delta _{\mathrm {F}}$ is included in $\Delta _N$, in cases where $\Delta _{\mathrm {F}}$ > 0 and $\Delta _{\mathrm {F}}$ < 0, different magnon gain $\mathcal {G}$ can be obtained, enabling the realization of nonreciprocal magnon laser. We define stimulated emission of magnon number is $M=\exp {[2(\mathcal {G}-\gamma _\textrm {m}/2)/(\gamma _\textrm {m}/2)]}$, which characterizes the performance of the magnon laser. It indicates the occurrence of magnon amplification when $M=1$, thus the threshold condition for magnon laser is $\mathcal {G}=\gamma _\textrm {m}/2$. From the threshold condition $\mathcal {G}=\gamma _\textrm {m}/2$, the threshold power of magnon laser can be derived as

$$P_\textrm{th}=\frac{-\frac{1}{4} \hbar\mathcal{A} \gamma_\textrm{m}\omega_\textrm{d}\left(\mathcal{B}+\left(\gamma^2+4 \tilde{\Delta}^2\right)\mathcal{A}-\mathcal{C} \phi_0+4\mathcal{A} \phi_0^2\right) \left({\gamma^{\prime}}^2+4\left(\Delta_\omega-\omega_\textrm{m}\right)^2\right)}{\left(2 \textit{G}^2 J\textit{S} \kappa \gamma^{\prime}\left(\tilde{\Delta}\left(3 J^2+4 \tilde{\Delta}^2+\kappa^2\right)-\mathcal{A} \phi_0\right)\right)}$$
where $\mathcal {A}=\left (4 \tilde {\Delta }^2+\kappa ^2\right )$, $\mathcal {B}=J^4-2 J^2\left (4 \tilde {\Delta }^2+\gamma \kappa \right )$ and $\mathcal {C}=8 \tilde {\Delta }\left (-J^2+4 \tilde {\Delta }^2+\kappa ^2\right )$.

3. Analysis and discussion

In above discussion, we have already determined the position of the EP and found the optimal coupling strength $J$. Therefore, we can research the maximum achievable gain and nonreciprocal effects of the system. We plot the magnon gain $\lg {(\mathcal {G})}$ as a function of normalized detuning $\Delta /\omega _\textrm {m}$, as shown in Fig. 3(a). It can be observed that when the fixed resonators case, i.e., $\Delta _{\mathrm {F}}=0$, the system can achieve maximum gain at $\Delta$ = $0.5~\omega _\textrm {m}$, and the maximum gain can reach magnitudes of $10^9$. In this case, regardless of the direction from which the pumping is input into the system, the resulting magnon gain $\mathcal {G}$ remains the same, indicating that the system is reciprocal at this point. On the other hand, when $\Delta _{\mathrm {F}}>0$ ($\Delta _{\mathrm {F}}<0$), it causes the maximum value of $\mathcal {G}$ to shift to the left (right), this results in directional correlation of our system. By selecting an appropriate detuning, it is possible to enhance the magnon gain when driving the resonator from one direction, while suppressing the magnon gain when driving the resonator from the opposite direction. This indicates that the system is nonreciprocal under these conditions. In order to verify that the system produces nonreciprocity, we plot the steady-state photon numbers $\left |a_{1s}\right |^2$ of the active optical WGM $a_1$ as functions of normalized detuning in Fig. 3(b) and the steady-state photon numbers $\left |a_{2s}\right |^2$ of the passive optical WGM $a_2$ as functions of normalized detuning in Fig. 3(c), respectively. Clearly can be seen, in different driving directions, the position of maximum steady-state photon number in the two spinning resonators experience a frequency shift. Therefore, by selecting specific detunings, significant differences in the steady-state photon number occur for two driving directions. These differences in steady-state photon number lead to variations in the magnetic dipole interaction strength. Thus, the magnon gain exhibits a substantial difference when the system is driven in opposite directions.

 figure: Fig. 3.

Fig. 3. (a) The magnon gain $\lg {(\mathcal {G})}$, (b) steady-state photon number $\left |a_{1s}\right |^2$ of active WGM mode, and (c) steady-state photon number $\left |a_{2s}\right |^2$ of passive WGM mode as a function of the normalized detuning $\Delta /\omega _\textrm {m}$. The parameters are used $\kappa /2\pi =15$ MHz, $\gamma /2\pi =0.1$ MHz, $G=1$ Hz [54,55], $S=1\times 10^{10}$, $\omega _\textrm {m}/2\pi =1$ GHz, $\Delta _{\mathrm {F}}=0.1\omega _m$, $\Delta \omega =\omega _\textrm {m}$, $\kappa /\gamma =0.8$, $\omega _\textrm {d}=193$ THz, $\gamma _\textrm {m}/2\pi =0.1$ MHz, $J=\sqrt {(\kappa +\gamma )^2/16+\omega _{\textrm {m}}^2/4}\approx \pi$ GHz [35,47,53,56].

Download Full Size | PDF

To better explain the influence of resonators rotation on magnon lasers, we introduce the isolation parameter $\mathfrak {R}$ = $10 \log _{10} [\mathcal {G}\left (\Delta _{\mathrm {F}}<0\right )]/[\mathcal {G}\left (\Delta _{\mathrm {F}}>0\right )]$. For the fixed resonators case, implying a reciprocal system, the isolation parameter $\mathfrak {R}=0$, however when the isolation parameter $\mathfrak {R}\neq 0$, it signifies the emergence of nonreciprocity in the system. Consequently, in Fig. 4, we plot the isolation parameter as a function of normalized detuning $\Delta /\omega _{\textrm {m}}$ and the Fizeau shift $\Delta _{\mathrm {F}}$. As the Fizeau shift $\Delta _{\mathrm {F}}$ increases, the nonreciprocity of the spin resonator becomes more obvious. By selecting an appropriate normalized detuning, we can achieve a pronounced nonreciprocal magnon laser phenomenon. Therefore, adjusting the driving direction of the resonator can be used as a switch to activate or deactivate the laser. This is helpful to understand and control of nonreciprocal behavior in magnon lasers, and it has significant implications for their practical applications in physics research.

 figure: Fig. 4.

Fig. 4. Isolation parameter $\mathfrak {R}$ versus the normalized detuning $\Delta /\omega _\textrm {m}$ and the Fizeau shift $\Delta _{\mathrm {F}}$ with the input power $P_{\textrm {in}}=2$ mW. The other parameters are the same as those in Fig. 3

Download Full Size | PDF

We further demonstrate the nonreciprocal characteristics of the system and the advantages of $\mathcal {PT}$-symmetric systems, which provide attractive opportunities for our system to generate one-way magnon laser across a range of input power while achieving ultra-low threshold power. Figure 5 shows the stimulated emitted magnon number $M$ versus the input power $P_{\textrm {in}}$ in the presence of $\Delta _{\mathrm {F}}<0$, $\Delta _{\mathrm {F}}=0$ and $\Delta _{\mathrm {F}}>0$ with different values of the active optical WGM gain coefficient $\kappa$. The shaded part of the figure indicates that magnon laser is below the threshold condition $M=1$. Based on previous research, it is known that laser generation occurs only when the system reaches the threshold condition, and the input power at the threshold condition is called the threshold power, and the three colored dots denote the threshold power $P_{\textrm {th}}$, respectively. We find that the stimulated emission magnon number $M$ exhibits strong nonlinearity with the input power and is directional dependent. When $\Delta _{\mathrm {F}}>0$, i.e., pumping light is input from the left, the threshold power $P_{\textrm {th}}$ is significantly lower than in the case of the opposite input direction and under the fixed resonators case. Taking Fig. 5(a) as an example, in the case of fixed resonator, the threshold power of the system $P_{\textrm {th}}\approx 17$ $\mu \mathrm {W}$. For $\Delta _{\mathrm {F}}>0$, the threshold power $P_{\textrm {th}}$ can be reduced to $2.3$ $\mu \mathrm {W}$, which is attributed to the enhancement of the magnon gain. For the case of $\Delta _{\mathrm {F}}<0$, the threshold power is as high as $49.2$ $\mu \mathrm {W}$. It is because the suppression of magnon gain at $\Delta = 0.34$ $\omega _\textrm {m}$, hence requiring a higher input power. Therefore, $\Delta _{\mathrm {F}}>0$ produces magnon laser when the input power is greater than $2.3$ $\mu \mathrm {W}$ and less than $49.2$ $\mu \mathrm {W}$, while when $\Delta _{\mathrm {F}}<0$ no magnon laser occurs at all because the threshold power is not reached. A one-way magnon laser is realized in the input power range. Figures 5(b)-(d) show the effects on the system threshold power as the optical gain-loss ratio increases to 0.85, 0.9, 0.95, respectively. Obviously, as the gain-loss ratio gradually approaches 1, the system can reach ultra-low threshold power, and even for $\Delta _{\mathrm {F}}>0$, the system can approach zero threshold power. This is because for a $\mathcal {PT}$-symmetric system, the dissipation $\gamma _{\pm }$ of the system can approach 0 when the gain-loss ratio is close to 1, so the energy exchange between the supermodes has a very low internal loss and the gain can be effectively introduced, thus achieving ultra-low threshold power.

 figure: Fig. 5.

Fig. 5. The stimulated emitted magnon number $M$ as a function of the driving power $P_{\textrm {in}}$ at the detuning $\Delta /\omega _\textrm {m}$ = 0.34 in the context of gain-loss ratio (a) $\kappa /\gamma =0.8$ (b) $\kappa /\gamma =0.85$ (c) $\kappa /\gamma =0.9$ (d) $\kappa /\gamma =0.95$. The other parameters are the same as those in Fig. 3.

Download Full Size | PDF

4. Conclusions

In conclusion, we have proposed the nonreciprocal $\mathcal {PT}$-symmetric magnon laser in a $\mathcal {PT}$-symmetric cavity optomagnonical system consists of active and passive optical spinning resonators as well as a YIG sphere. By introducing the Fizeau light-dragging induced by the rotation of resonators, we observe the significant variations in the steady-state photon populations of active optical WGM and passive optical WGM as well as magnon gain. It is verified that the trend of the magnon gain is the same as the trend of the photon populations number in both resonators. We also show that the Fizeau shift can further reduce the threshold power of the system, allowing the system to approach 0 threshold power without reaching gain-loss balance, and achieve higher isolation rates within experimentally feasible parameter ranges. Furthermore, by controlling the rotational speed of the spinning resonators and driving direction, the magnon gain and threshold power can be flexible adjusted. Notably, we successfully demonstrate a one-way magnon laser can be realized across a range of input power. Thus a nonreciprocal $\mathcal {PT}$-symmetric magnon laser can be achieved. This provides a flexible method for manipulating non-Hermitian cavity magnon systems and holds potential applications in various spin-electronics devices.

Funding

National Natural Science Foundation of China (62101479, 162071412, 12074330, 12375020).

Disclosures

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

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. Ö. O. Soykal and M. Flatté, “Strong field interactions between a nanomagnet and a photonic cavity,” Phys. Rev. Lett. 104(7), 077202 (2010). [CrossRef]  

2. Z. X. Liu, C. You, B. Wang, et al., “Phase-mediated magnon chaos-order transition in cavity optomagnonics,” Opt. Lett. 44(3), 507–510 (2019). [CrossRef]  

3. X Zhang, “A Review of Common Materials for Hybrid Quantum Magnonics,” Mater. Today Electron. 5, 100044 (2023). [CrossRef]  

4. W. Zhang, D. Y. Wang, C. H. Bai, et al., “Generation and transfer of squeezed states in a cavity magnomechanical system by two-tone microwave fields,” Opt. Express 29(8), 11773–11783 (2021). [CrossRef]  

5. K Wang, Y. P. Gao, R Jiao, et al., “Recent progress on optomagnetic coupling and optical manipulation based on cavity-optomagnonics,” Front. Phys. 17(4), 42201 (2022). [CrossRef]  

6. W. L. Xu, Y. P. Gao, T. J. Wang, et al., “Magnon-induced optical high-order sideband generation in hybrid atom-cavity optomagnonical system,” Opt. Express 28(15), 22334–22344 (2020). [CrossRef]  

7. W. Zhang, T. Wang, X. Han, et al., “Quantum entanglement and one-way steering in a cavity magnomechanical system via a squeezed vacuum field,” Opt. Express 30(7), 10969–10980 (2022). [CrossRef]  

8. X. Zhang, C. L. Zou, L. Jiang, et al., “Strongly coupled magnons and cavity microwave photons,” Phys. Rev. Lett. 113(15), 156401 (2014). [CrossRef]  

9. J. Cheng, Y. M. Liu, X. Yi, et al., “Generation and enhancement of mechanical squeezing in a hybrid cavity magnomechanical system,” Annalen Phys. 534(5), 2100493 (2022). [CrossRef]  

10. M. Goryachev, W. G. Farr, D. L. Creedon, et al., “High-cooperativity cavity qed with magnons at microwave frequencies,” Phys. Rev. Appl. 2(5), 054002 (2014). [CrossRef]  

11. T. X. Lu, X Xiao, L. S. Chen, et al., “Magnon-squeezing-enhanced slow light and second-order sideband in cavity magnomechanics,” Phys. Rev. A 107(6), 063714 (2023). [CrossRef]  

12. D. Zhang, X. M. Wang, T. F. Li, et al., “Cavity quantum electrodynamics with ferromagnetic magnons in a small yttrium-iron-garnet sphere,” NPJ Quantum Inf. 1(1), 15014 (2015). [CrossRef]  

13. Y. Tabuchi, S. Ishino, T. Ishikawa, et al., “Hybridizing ferromagnetic magnons and microwave photons in the quantum limit,” Phys. Rev. Lett. 113(8), 083603 (2014). [CrossRef]  

14. X. Zhang, C. L. Zou, N. Zhu, et al., “Magnon dark modes and gradient memory,” Nat. Commun. 6(1), 8914 (2015). [CrossRef]  

15. Y. P. Gao, C Cao, Y. W. Duan, et al., “Magnons scattering induced photonic chaos in the optomagnonic resonators,” Nanophotonics 9(7), 1953–1961 (2020). [CrossRef]  

16. D. Zhang, X. Q. Luo, Y. P. Wang, et al., “Observation of the exceptional point in cavity magnon-polaritons,” Nat. Commun. 8(1), 1368 (2017). [CrossRef]  

17. X. Zhang, K. Ding, X. Zhou, et al., “Experimental observation of an exceptional surface in synthetic dimensions with magnon polaritons,” Phys. Rev. Lett. 123(23), 237202 (2019). [CrossRef]  

18. J. Zhao, Y. Liu, L. Wu, et al., “Observation of anti-p t-symmetry phase transition in the magnon-cavity-magnon coupled system,” Phys. Rev. Appl. 13(1), 014053 (2020). [CrossRef]  

19. M. Harder, Y. Yang, B. Yao, et al., “Level attraction due to dissipative magnon-photon coupling,” Phys. Rev. Lett. 121(13), 137203 (2018). [CrossRef]  

20. V. L. Grigoryan, K. Shen, and K. Xia, “Synchronized spin-photon coupling in a microwave cavity,” Phys. Rev. B 98(2), 024406 (2018). [CrossRef]  

21. W. Yu, J. Wang, H. Yuan, et al., “Prediction of attractive level crossing via a dissipative mode,” Phys. Rev. Lett. 123(22), 227201 (2019). [CrossRef]  

22. S. Y. Guan, H. F. Wang, and X. Yi, “Cooperative-effect-induced one-way steering in open cavity magnonics,” NPJ Quantum Inf. 8(1), 102 (2022). [CrossRef]  

23. J. Cheng, Y. M. Liu, H. F. Wang, et al., “Entanglement and asymmetric steering between distant magnon and mechanical modes in an optomagnonic-mechanical system,” Annalen Phys. 534(12), 2200315 (2022). [CrossRef]  

24. Z. Zhang, Y. Zhang, J. Sheng, et al., “Observation of parity-time symmetry in optically induced atomic lattices,” Phys. Rev. Lett. 117(12), 123601 (2016). [CrossRef]  

25. B. Wang, Z. X. Liu, X. Jia, et al., “Polarization-based control of phonon laser action in a Parity Time-symmetric optomechanical system,” Commun. Phys. 1(1), 43 (2018). [CrossRef]  

26. T. Wang, W. Zhang, J. Cao, et al., “Exceptional-point-engineered phonon laser in a cavity magnomechanical system,” New J. Phys. 25(8), 083045 (2023). [CrossRef]  

27. H. Jing, S. Özdemir, X. Y. Lü, et al., “$\mathcal {PT}$-symmetric phonon laser,” Phys. Rev. Lett. 113(5), 053604 (2014). [CrossRef]  

28. X. Y. Lü, H. Jing, J. Y. Ma, et al., “$\mathcal {PT}$-symmetry-breaking chaos in optomechanics,” Phys. Rev. Lett. 114(25), 253601 (2015). [CrossRef]  

29. H. Zhang, R. Huang, S. D. Zhang, et al., “Breaking anti-$\mathcal {PT}$ symmetry by spinning a resonator,” Nano Lett. 20(10), 7594–7599 (2020). [CrossRef]  

30. W. L. Xu, X. F. Liu, Y. Sun, et al., “Magnon-induced chaos in an optical $\mathcal {PT}$-symmetric resonator,” Phys. Rev. E 101(1), 012205 (2020). [CrossRef]  

31. G. Q. Zhang, Z. Chen, D. Xu, et al., “Exceptional point and cross-relaxation effect in a hybrid quantum system,” PRX Quantum 2(2), 020307 (2021). [CrossRef]  

32. L. Wang, Z. X. Yang, Y. M. Liu, et al., “Magnon blockade in a $\mathcal {PT}$-symmetric-like cavity magnomechanical system,” Annalen Phys. 532(7), 2000028 (2020). [CrossRef]  

33. B. Wang, Z. X. Liu, C. Kong, et al., “Magnon-induced transparency and amplification in $\mathcal {PT}$-symmetric cavity-magnon system,” Opt. Express 26(16), 20248–20257 (2018). [CrossRef]  

34. A. Osada, R. Hisatomi, A. Noguchi, et al., “Cavity optomagnonics with spin-orbit coupled photons,” Phys. Rev. Lett. 116(22), 223601 (2016). [CrossRef]  

35. B. Wang, X. Jia, X. H. Lu, et al., “$\mathcal {PT}$-symmetric magnon laser in cavity optomagnonics,” Phys. Rev. A 105(5), 053705 (2022). [CrossRef]  

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

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

38. Z. Shen, Y. L. Zhang, Y. Chen, et al., “Experimental realization of optomechanically induced non-reciprocity,” Nat. Photonics 10(10), 657–661 (2016). [CrossRef]  

39. N. Bender, S. Factor, J. D. Bodyfelt, et al., “Observation of asymmetric transport in structures with active nonlinearities,” Phys. Rev. Lett. 110(23), 234101 (2013). [CrossRef]  

40. L. Chang, X. Jiang, S. Hua, et al., “Parity–time symmetry and variable optical isolation in active–passive-coupled microresonators,” Nat. Photonics 8(7), 524–529 (2014). [CrossRef]  

41. R. Huang, A. Miranowicz, J. Q. Liao, et al., “Nonreciprocal photon blockade,” Phys. Rev. Lett. 121(15), 153601 (2018). [CrossRef]  

42. B. Li, R. Huang, X. Xu, et al., “Nonreciprocal unconventional photon blockade in a spinning optomechanical system,” Photonics Res. 7(6), 630–641 (2019). [CrossRef]  

43. Y. M. Liu, J. Cheng, H. F. Wang, et al., “Nonreciprocal photon blockade in a spinning optomechanical system with nonreciprocal coupling,” Opt. Express 31(8), 12847–12864 (2023). [CrossRef]  

44. W. Zhang, T. Wang, S. Liu, et al., “Nonreciprocal photon blockade in a spinning resonator coupled to two two-level atoms,” Sci. China Phys. Mech. Astron. 66(4), 240313 (2023). [CrossRef]  

45. Y. Jiang, S. Maayani, T. Carmon, et al., “Nonreciprocal phonon laser,” Phys. Rev. Appl. 10(6), 064037 (2018). [CrossRef]  

46. Y. Xu, J. Y. Liu, W. Liu, et al., “Nonreciprocal phonon laser in a spinning microwave magnomechanical system,” Phys. Rev. A 103(5), 053501 (2021). [CrossRef]  

47. Z. X. Liu and H. Xiong, “Magnon laser based on brillouin light scattering,” Opt. Lett. 45(19), 5452–5455 (2020). [CrossRef]  

48. Y. J. Xu and J. Song, “Nonreciprocal magnon laser,” Opt. Lett. 46(20), 5276–5279 (2021). [CrossRef]  

49. K. W. Huang, Y. Wu, and L. G. Si, “Parametric-amplification-induced nonreciprocal magnon laser,” Opt. Lett. 47(13), 3311–3314 (2022). [CrossRef]  

50. X. W. He, Z. Y. Wang, X Han, et al., “Parametrically amplified nonreciprocal magnon laser in a hybrid cavity optomagnonical system,” Opt. Express 31(26), 43506–43517 (2023). [CrossRef]  

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

52. B. Peng, S. K. Özdemir, F. Lei, et al., “Parity–time-symmetric whispering-gallery microcavities,” Nat. Phys. 10(5), 394–398 (2014). [CrossRef]  

53. S. V. Kusminskiy, H. X. Tang, and F. Marquardt, “Coupled spin-light dynamics in cavity optomagnonics,” Phys. Rev. A 94(3), 033821 (2016). [CrossRef]  

54. J. Graf, S. Sharma, H. Huebl, et al., “Design of an optomagnonic crystal: Towards optimal magnon-photon mode matching at the microscale,” Phys. Rev. Res. 3(1), 013277 (2021). [CrossRef]  

55. N. Zhu, X. Zhang, X. Han, et al., “Waveguide cavity optomagnonics for microwave-to-optics conversion,” Optica 7(10), 1291–1297 (2020). [CrossRef]  

56. X. Zhang, N. Zhu, C. L. Zou, et al., “Optomagnonic whispering gallery microresonators,” Phys. Rev. Lett. 117(12), 123605 (2016). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic diagram of a nonreciprocal $\mathcal {PT}$-symmetric cavity optomagnonical system. The ferrimagnet YIG sphere supports optical WGM $a_{2}$ and magnon mode $m$ and optical WGM $a_{2}$ couples to active optical WGM $a_{1}$. The input light evanescent couples to the active optical WGM via a tapered fiber. Driving the system from the left ($\Delta _{\mathrm {F}}>0$) enhances the magnon laser, while driving from the right ($\Delta _{\mathrm {F}}<0$) suppresses or even blocks magnon laser emission.
Fig. 2.
Fig. 2. (a) Eigenfrequencies splitting of supermodes $(\omega _{\pm }-\tilde {\Delta })/\gamma$ and (b) system dissipation of the eigenmodes $\gamma _{\pm }/\gamma$ vary with the optical tunneling rate $J/\gamma$ under the condition of $\kappa /\gamma$ = 1.
Fig. 3.
Fig. 3. (a) The magnon gain $\lg {(\mathcal {G})}$, (b) steady-state photon number $\left |a_{1s}\right |^2$ of active WGM mode, and (c) steady-state photon number $\left |a_{2s}\right |^2$ of passive WGM mode as a function of the normalized detuning $\Delta /\omega _\textrm {m}$. The parameters are used $\kappa /2\pi =15$ MHz, $\gamma /2\pi =0.1$ MHz, $G=1$ Hz [54,55], $S=1\times 10^{10}$, $\omega _\textrm {m}/2\pi =1$ GHz, $\Delta _{\mathrm {F}}=0.1\omega _m$, $\Delta \omega =\omega _\textrm {m}$, $\kappa /\gamma =0.8$, $\omega _\textrm {d}=193$ THz, $\gamma _\textrm {m}/2\pi =0.1$ MHz, $J=\sqrt {(\kappa +\gamma )^2/16+\omega _{\textrm {m}}^2/4}\approx \pi$ GHz [35,47,53,56].
Fig. 4.
Fig. 4. Isolation parameter $\mathfrak {R}$ versus the normalized detuning $\Delta /\omega _\textrm {m}$ and the Fizeau shift $\Delta _{\mathrm {F}}$ with the input power $P_{\textrm {in}}=2$ mW. The other parameters are the same as those in Fig. 3
Fig. 5.
Fig. 5. The stimulated emitted magnon number $M$ as a function of the driving power $P_{\textrm {in}}$ at the detuning $\Delta /\omega _\textrm {m}$ = 0.34 in the context of gain-loss ratio (a) $\kappa /\gamma =0.8$ (b) $\kappa /\gamma =0.85$ (c) $\kappa /\gamma =0.9$ (d) $\kappa /\gamma =0.95$. The other parameters are the same as those in Fig. 3.

Equations (15)

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

Δ F = ± Ω n r ω c c ( 1 1 n 2 λ n d n d λ ) ,
H =   H 0 + H int + H dr , H 0 =   ( Δ Δ F ) ( a 1 a 1 + a 2 a 2 ) + ω m m m , H int =   J ( a 1 a 2 + a 2 a 1 ) + G S / 2 a 2 a 2 ( m + m ) , H dr =   i κ / 2 ε d ( a 1 a 1 ) ,
a ˙ 1 =   ( i Δ + i Δ F + κ / 2 ) a 1 i J a 2 + κ / 2 ε d , a ˙ 2 =   ( i Δ + i Δ F γ / 2 ) a 2 i J a 1 i G S / 2 a 2 ( m + m ) , m ˙ =   i ω m m γ m / 2 m i G S / 2 a 2 a 2 ,
a 1 s =   i J a 2 s κ / 2 ε d i Δ ~ + κ / 2 , a 2 s =   4 J κ / 2 ε d i ( 4 Δ ~ 2 + J 2 κ γ ) + 2 ( γ + 2 i ϕ 0 κ ) Δ ~ + 2 κ ϕ 0 , m s =   i G S / 2 | a 2 s | 2 ( i ω m + γ m / 2 ) ,
H non = ( Δ ~ + i κ / 2 ) a 1 a 1 + ( Δ ~ i γ / 2 ) a 2 a 2 + J ( a 1 a 2 + a 2 a 1 ) ,
H = ω + Ψ + Ψ + + ω Ψ Ψ + ω m m m G S / 2 2 ( Ψ + Ψ m + m Ψ Ψ + ) ,
Ψ ˙ + =   ( i ω + + κ / 2 ) Ψ + + i G S / 2 2 Ψ m , Ψ ˙ =   ( i ω γ / 2 ) Ψ + i G S / 2 2 Ψ + m , m ˙ =   ( i ω m γ m / 2 ) m + i G S / 2 2 Ψ Ψ + .
m ˙ = ( i ω m γ m / 2 ) m + i G S / 2 / 2 p , p ˙ = ( i Δ ω γ / 2 ) p i G S / 2 Δ N / 2 m ,
m ~ = U m U = m e i ω m t , p ~ = U p U = p e i Δ ω t .
m ~ ˙ = γ m / 2 m ~ + i G S / 2 / 2 Φ ~ e i ( ω m Δ ω ) t ,
Φ ~ ˙ = γ / 2 Φ ~ i G S / 2 Δ N / 2 m ~ e i ( Δ ω ω m ) t .
p ~ = i G S / 2 Δ N / 2 e i ( Δ ω ω m ) t i ( Δ ω ω m ) + γ 2 m ~ .
m ~ = exp [ ( γ m / 2 + G i Ω m ) t ] .
G = G 2 S Δ N γ / 16 ( Δ ω ω m ) 2 + ( γ 2 ) 2 ,
P th = 1 4 A γ m ω d ( B + ( γ 2 + 4 Δ ~ 2 ) A C ϕ 0 + 4 A ϕ 0 2 ) ( γ 2 + 4 ( Δ ω ω m ) 2 ) ( 2 G 2 J S κ γ ( Δ ~ ( 3 J 2 + 4 Δ ~ 2 + κ 2 ) A ϕ 0 ) )
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.