Chinese Physics Letters, 2023, Vol. 40, No. 10, Article code 104201 Nonreciprocal Phonon Laser in an Asymmetric Cavity with an Atomic Ensemble Kai-Wei Huang (黄凯伟), Xin Wang (王鑫), Qing-Yang Qiu (邱庆洋), Long Wu (吴龙), and Hao Xiong (熊豪)* Affiliations School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China Received 25 June 2023; accepted manuscript online 7 September 2023; published online 3 October 2023 *Corresponding author. Email: haoxiong1217@gmail.com Citation Text: Huang K W, Wang X, Qiu Q Y et al. 2023 Chin. Phys. Lett. 40 104201    Abstract Phonon lasers, as a novel kind of lasers for generating coherent sound oscillation, has attracted extensive attention. Here, we theoretically propose a nonreciprocal phonon laser in a hybrid optomechanical system, which is composed of an asymmetric Fabry–Pérot cavity, an ensemble of $N$ identical two-level atoms, and a mechanical oscillator. The effective driving amplitude related to driving direction leads to an obvious difference in mechanical gain and threshold power, bringing about a nonreciprocal phonon laser. In addition, the dependence of the phonon laser on the atomic parameters is also discussed, including the decay rate of the atoms and the coupling strength between the atoms and the cavity field, which provides an additional degree of freedom to control the phonon laser action. Our work provides a path to realizing a phonon laser in an atoms-cavity optomechanical system and may aid the manufacture of directional coherent phonon sources.
cpl-40-10-104201-fig1.png
cpl-40-10-104201-fig2.png
cpl-40-10-104201-fig3.png
cpl-40-10-104201-fig4.png
DOI:10.1088/0256-307X/40/10/104201 © 2023 Chinese Physics Society Article Text A Fabry–Pérot cavity is composed of two fixed mirrors at both ends of the cavity, which can capture light to enhance the light-matter interaction. Generally, a movable mirror (i.e., cantilever tip)[1,2] is introduced into the cavity to resonate as a harmonic near the balance position,[3] realizing the interaction between light and matter.[4] Due to its great application potential in cooling of mechanical oscillators,[5-7] highly sensitive measurements,[8-12] and gravitational wave detectors,[13,14] this setup, cavity optomechanics (COM),[15,16] has aroused great interests in recent years. This novel mechanism enriches the control methods of photons, and allows the investigation of some fascinating phenomena in the classical and quantum regimes,[17-27] such as optomechanically induced transparency, high-order sideband generation, photon blockade, and entanglement. Further, phonon lasers have been studied experimentally and theoretically[28-33] by manipulating phonons in COM systems. Compared to optical lasers, a phonon laser[34] has the following advantages: the phonon cavity is easy to tune, the speed of a sound wave is significantly slower than that of light, and the phonon is not affected by radiation loss in a vacuum. Thus, phonon lasers can achieve high energy concentration, and perform high-precision non-destructive measurements. In addition, the scheme of phonon lasers can also be achieved in a quantum-dot system.[35,36] A phonon laser generating a coherent phonon source provides an important platform in functional phononic devices[37] and quantum acoustics.[38-40] In parallel, the nonreciprocal devices that break the physical symmetry enable light to propagate from one direction rather than the other, which plays an extremely important role in wide applications of circulators,[41,42] optical isolators,[43,44] and metamaterials.[45,46] Traditional nonreciprocal devices based on the magneto-optical Faraday effect[47] not only need strong magnetic field and bulky materials, but are also difficult for on-chip integration. In order to overcome these shortcomings, various schemes of nonreciprocal optical devices have been implemented in non-Hermitian systems[48,49] and Kerr resonators,[50,51] as well as time modulation for photonic systems.[52] In particular, many nonreciprocal phenomena have been reported in spinning resonators based on the Sagnac effect,[22,53-57] such as photon blockade, phonon lasers, entanglement, chaos, and sideband responses. Specifically, by coupling atoms with an asymmetric cavity field, optical nonreciprocity transmission has been demonstrated.[58-60] However, as far as we know, the nonreciprocal phonon laser in an asymmetric cavity is rarely considered and remains to be investigated. Inspired by the advantages of the asymmetric cavity system, we study a hybrid COM system, in which our system is composed of an asymmetric Fabry–Pérot cavity, a two-level atomic ensemble, and a mechanical oscillator to realize a nonreciprocal phonon laser. The hybrid COM system offers the phonon analogue of a two-level laser through phonon mediated optical transitions. Then, the optical inversion generates coherent mechanical gain at the breathing mode, resulting in a phonon laser above the threshold power $P_{\rm{th}}$.[28] The gain factor of the phonon laser increases from one driving direction and suppresses from the opposite driving direction. The fundamental reason for this is that the attenuations of the asymmetric cavity walls are different, resulting in different driving amplitudes at the same driving power. We find that the isolation parameter linearly changes with the attenuation of the cavity walls and can reach about 24.2 dB. In addition, we introduce an ensemble of atoms to replace one of the cavities in the previously studied phonon laser system,[28,29,54] which provides a new method for controlling the phonon laser. Compared to the previous scheme in a spinning resonator,[54] our system is more robust, since it does not need to maintain constant rotational speed and to consider additional rotational kinetic energy. Here, our results provide a method to adjust the phonon laser by the driving direction and the atomic parameters, and offers a guideline for further experiments. Our scheme may provide applications for the atoms-COM system in quantum information processing and nonreciprocal devices.
cpl-40-10-104201-fig1.png
Fig. 1. (a) A schematic diagram of our hybrid optomechanical system. The asymmetric Fabry–Pérot cavity contains an ensemble of $N$ identical two-level atoms and a movable membrane, in which the atoms and the mechanical oscillators interact with cavity modes. Here cavity decay rate $\kappa_1$ ($\kappa_2$) denotes the loss of cavity mode via $M_1$ ($M_2$); $a_{\rm{in}}$ ($b_{\rm{in}}$) is forward (backward) input with the same power. Levels $|g\rangle$ and $|e\rangle$ represent the ground and excited states of the atom. (b) Scheme depicting the corresponding optical supermodes coupled by a phonon, as well as the equivalent two-level phonon laser energy-level diagram.
As shown in Fig. 1(a), we present a hybrid optomechanical system including the asymmetrical Fabry–Pérot cavity, the membrane, and the two-level atoms. A movable membrane is regarded as a quantum mechanical harmonic oscillator with frequency $\omega_{\rm m}$ and interacts with the cavity field through the radiation pressure. The two-level atomic ensemble with frequency $\omega_{\rm a}$ is confined by the magneto-optical trap in the cavity and couples to the cavity mode via the dipole interaction.[59] Such a setup of our system can be realized in a SiN membrane on a silicon chip containing a $^{87}$Rb atom vapor with buffer gas.[61,62] The cavity mode with frequency $\omega_{\rm c}$ is driven by the pump field with frequency $\omega_{\scriptscriptstyle{\rm L}}$. The cavity decay rates $\kappa_1$ and $\kappa_2$ of the cavity walls $\rm{M_1}$ and $\rm{M_2}$ meet the conditions $\kappa_i = -c\cdot\ln (R_i/2L)$, where $R_i$ is the reflectivity of cavity wall ${\rm{M}}_{i}$ ($i=1,\,2$), $L$ is the effective length of the optical cavity, and $c$ is the speed of light in vacuum.[59,60] Generally, the decay rates $\kappa_1$ and $\kappa_2$ of the cavity walls are unequal, thus breaking the spatial symmetry. Only when $\kappa_1 = \kappa_2 = \kappa_0$, the cavity symmetry is satisfied, where $\kappa_0 = (\kappa_1+\kappa_2)/2$ is defined as the average loss rate of the cavity. The total decay of the cavity is $\kappa_{\rm c}=\kappa_1+\kappa_2+\kappa_{\rm loss}$, where the remaining dissipation rate of the cavity mode $\kappa_{\rm loss}$ can be neglected in the high-quality cavity ($\kappa_{\rm loss}/\kappa_{\rm c}\ll1$). Then, the Hamiltonian of our system can be written as ($\hbar=1$) \begin{align} &H=H_0+H_{\rm{int}}+H_{\rm{dr}},\notag\\ &H_0=\omega_{\rm c}c^†c+\omega_{\rm m}b^†b+\sum\limits^{N}_{i=1}\omega_{\rm a}\sigma^{(i)}_{\rm ee},\notag\\ &H_{\rm{int}}=gc^†c(b^†+b)+\lambda\sum\limits^{N}_{i=1}(c\sigma^{(i)}_{\rm eg}+c^†\sigma^{(i)}_{\rm ge}),\notag\\ &H_{\rm{dr}}=i(\varepsilon_{\scriptscriptstyle{\rm L}}c^†e^{-i\omega_{\scriptscriptstyle{\rm L}}t}-\varepsilon_{\scriptscriptstyle{\rm L}}^{\ast}ce^{i\omega_{\scriptscriptstyle{\rm L}} t}). \tag {1} \end{align} Here, $H_0$ is the free Hamiltonian of the system, $c$ ($c^†$) and $b$ ($b^†$) are the annihilation (creation) operators of the photon and phonon, respectively, and the third term represents the population operator of the $i$th atom. In addition, when $e\neq g$, $\sigma^{(i)}_{\rm eg}$ is the electronic projection operator. $H_{\rm{int}}$ is the interaction Hamiltonian. The first term describes the optomechanical coupling between the photon and mechanical mode with the coupling strength $g$. The second term describes the dipole interactions between the photon and two-level atoms with the coupling strength $\lambda=\mu\sqrt{\omega_{\rm c}/2\hbar V\varepsilon_0}$, where $\mu$, $V$, and $\varepsilon_0$ are the dipole moment, the cavity volume, and the vacuum permittivity, respectively. $H_{\rm{dr}}$ denotes the driving field, where the effective driving strength is $\varepsilon_{\scriptscriptstyle{\rm L}} = \sqrt{2\kappa_i P_{\rm{in}}/\hbar\omega_{\scriptscriptstyle{\rm L}}}$ $(i=0,\,1,\,2)$.[59,60] For a fixed power $P_{\rm{in}}$, obviously the effective driving strength $\varepsilon_{\scriptscriptstyle{\rm L}}$ depends on the driving direction, as $\kappa_1$ and $\kappa_2$ are different. As shown in Fig. 1(a), we represent $a_{\rm{in}}$ as the case of the forward input and $b_{\rm{in}}$ as the case of backward input. We can acquire the gain factor of the phonon laser in the backward case by replacing $\kappa_2$ with $\kappa_1$. When the most atoms are initially prepared in the ground state, we can assume that these atoms exist in the cavity as a whole, which has been proved in previous studies.[63,64] Then, we can define a collective transition operator $a = \mathop{\rm{lim}}\limits_{N\rightarrow\infty}\sum^{N}_{i=1}(\lambda^{\ast}_i/J)|g\rangle_{ii}\langle e|$ fulfilling the bosonic commutation relation $[a,a^†] =1$, where $J = \lambda\sqrt{N}=\sqrt{\sum^{N}_{i=1}|\lambda_i|^2}$ is the total coupling strength between the atomic ensemble and the cavity field. The coupling parameter $J$ can be enhanced with the increase of atomic number $N$. In a frame rotating at frequency $\omega_{\scriptscriptstyle{\rm L}}$, the Hamiltonian of the system can be rewritten as \begin{align} H=\,&-\varDelta_{\rm c}c^†c-\varDelta_{\rm a}a^†a+\omega_{\rm m}b^†b+J(c^†a+ca^†)\notag\\ &+gc^†c(b^†+b)+i(\varepsilon_{\scriptscriptstyle{\rm L}}c^†-\varepsilon_{\scriptscriptstyle{\rm L}}^{\ast}c), \tag {2} \end{align} where $\varDelta_{\rm c}=\omega_{\scriptscriptstyle{\rm L}}-\omega_{\rm c}$ and $\varDelta_{\rm a}=\omega_{\scriptscriptstyle{\rm L}}-\omega_{\rm a}$ are the detunings of the cavity field frequency and the atomic transition frequency, respectively. The Heisenberg motion equations of this atoms-COM system are \begin{align} &\dot{c}=(i\varDelta_{\rm c}-\kappa_{\rm c})c-ig(b^†+b)c-iJa+\varepsilon_{\scriptscriptstyle{\rm L}},\notag\\ &\dot{a}=(i\varDelta_{\rm a}-\kappa_{\rm a})a-iJc,\notag\\ &\dot{b}=-(i\omega_{\rm m}+\gamma_{\rm b})b-igc^†c, \tag {3} \end{align} where $\kappa_{\rm a}$ and $\gamma_{\rm b}$ are the decay rate of the two-level atom and mechanical mode, respectively. The quantum noise term can be safely ignored under strong driving conditions, when we only focus on the average number actions.[65] For convenience, we assume that the atom and photon modes are resonant, that is, $\varDelta_{\rm c} = \varDelta_{\rm a} = \varDelta$. Then we can simply derive the steady-state solutions of this system as \begin{align} &c_{\rm s}=\dfrac{-(i\varDelta-\kappa_{\rm a})\varepsilon_{\scriptscriptstyle{\rm L}}}{[i\varDelta-ig(b^†+b)-\kappa_{\rm c}](i\varDelta-\kappa_{\rm a})+J^2},\notag\\ &a_{\rm s}=\dfrac{-iJ\varepsilon_{\scriptscriptstyle{\rm L}}}{[i\varDelta-ig(b^†+b)-\kappa_{\rm c}](i\varDelta-\kappa_{\rm a})+J^2},\notag\\ &b_{\rm s}=\dfrac{-ig|c_{\rm s}|^2}{i\omega_{\rm m}+\gamma_{\rm b}}. \tag {4} \end{align} Similar to optical lasers, the composite resonator can be used to realize the coherent emission of phonons through the inversion of two supermodes.[28] By introducing supermode operators $a_{\pm}=(a\pm c)/\sqrt{2}$, the Hamiltonian $H$ in Eq. (2) can be rewritten under the condition of the rotating-wave approximation as \begin{align} H=\,&\omega_{+}a^†_{+}a_{+}+\omega_{-}a^†_{-}a_{-}+\omega_{\rm m}b^†b\notag\\ &+\frac{g}{2}(a^†_{+}a_{-}b+a_{+}a^†_{-}b^†)\notag\\ &+\dfrac{i}{\sqrt{2}}[\varepsilon_{\scriptscriptstyle{\rm L}}(a^†_{+}+a^†_{-})-\rm{H.c.}], \tag {5} \end{align} where the supermode frequencies $\omega_{\pm}=-\varDelta\pm J$. The first interaction term in the Hamiltonian (5) describes the annihilation of a phonon and the optical mode transition from the red supermode to the blue supermode, while the second term describes the opposite process, as shown in Fig. 1(b). This can also be seen as stimulated Brillouin scattering, where a blue supermode is converted into a red supermode and a phonon with energy equal to the difference. In the supermode picture, the dynamical equations of the system can be obtained as \begin{align} &\dot{a}_{+}=-(i\omega_{+}+\kappa)a_{+}-\frac{i}{2}gba_{-}+\dfrac{\varepsilon_{\scriptscriptstyle{\rm L}}}{\sqrt{2}},\notag\\ &\dot{a}_{-}=-(i\omega_{-}+\kappa)a_{-}-\frac{i}{2}gb^†a_{+}+\dfrac{\varepsilon_{\scriptscriptstyle{\rm L}}}{\sqrt{2}},\notag\\ &\dot{b}=-(i\omega_{\rm m}+\gamma_{\rm b})b-\frac{i}{2}ga_{+}a^†_{-}, \tag {6} \end{align} with $\kappa=(\kappa_{\rm c}+\kappa_{\rm a})/2$. We set the ladder operator as $p=a^†_{-}a_{+}$ and the population difference $\delta n=a^†_{+}a_{+}-a^†_{-}a_{-}$ between the optical supermodes $a_{+}$ and $a_{-}$. Then, the motion equations of the system are rewritten as \begin{align} &\dot{b}=-(i\omega_{\rm m}+\gamma_{\rm b})b-\frac{i}{2}gp,\notag\\ &\dot{p}=-2(\kappa+iJ)p+\frac{i}{2}gb\delta n+\frac{1}{\sqrt{2}}(\varepsilon_{\scriptscriptstyle{\rm L}}^{\ast}a_{+}+\varepsilon_{\scriptscriptstyle{\rm L}}a^†_{-}). \tag {7} \end{align} Due to the fact that the supermode decay is more rapid than the decay of the mechanical oscillator, i.e., $\kappa\gg\gamma_{\rm b}$, compared with the variation of $a_{\pm}$ and $p$, $b$ can be seen as a constant. Thus, the supermode degree of freedom can be adiabatically eliminated. Then, we obtain the steady states of the system as \begin{align} &a_{+}=\dfrac{\varepsilon_{\scriptscriptstyle{\rm L}}(2i\omega_{-}+2\kappa-igb)}{2\sqrt{2}(\beta-2i\kappa\varDelta)},\notag\\ &a_{-}=\dfrac{\varepsilon_{\scriptscriptstyle{\rm L}}(2i\omega_{+}+2\kappa-igb^†)}{2\sqrt{2}(\beta-2i\kappa\varDelta)},\notag\\ &p=\dfrac{igb\delta n+\sqrt{2}(\varepsilon_{\scriptscriptstyle{\rm L}}^{\ast}a_{+}+\varepsilon_{\scriptscriptstyle{\rm L}}a^†_{-})}{2[2(\kappa+iJ)-i\omega_{\rm m}]}, \tag {8} \end{align} with $\beta=J^2-\varDelta^2+\kappa^2+g^2|b|^2/4$. The result for $b$ can be found by substituting Eq. (8) into Eq. (7) so that $\dot{b}=(-i\omega_{\rm m}-i\omega_{0}+G-\gamma_{\rm b})b+D$, with \begin{align} \omega_{0}=\,&\dfrac{g^2(2\,J-\omega_{\rm m})\delta n}{4(2\,J-\omega_{\rm m})^2+16\kappa^2}\notag\\ &+\dfrac{g^2|\varepsilon_{\scriptscriptstyle{\rm L}}|^2\kappa^2\varDelta}{[(2\,J-\omega_{\rm m})^2+4\kappa^2](\beta^2+4\varDelta^2\kappa^2)},\notag\\ D=\,&\dfrac{(\beta+2\varDelta^2)[2\kappa g(2\,J-\omega_{\rm m})-2ig\kappa^2]}{2[(2\,J-\omega_{\rm m})^2+4\kappa^2](\beta^2+4\varDelta^2\kappa^2)},\notag \end{align} and the mechanical gain of the phonon laser is $G=G_1+G_2$, with \begin{align} &G_1=\dfrac{g^2\kappa\delta n}{2[4\kappa^2+(2\,J-\omega_{\rm m})^2]},\notag\\ &G_2=\dfrac{|\varepsilon_{\scriptscriptstyle{\rm L}}|^2g^2(\omega_{\rm m}-2\,J)\varDelta\kappa}{2(\beta^2+4\kappa^2\varDelta^2)[4\kappa^2+(2\,J-\omega_{\rm m})^2]}, \tag {9} \end{align} where \begin{align} \delta n\approx\dfrac{2J\varDelta|\varepsilon_{\scriptscriptstyle{\rm L}}|^2}{\beta^2+4\kappa^2\varDelta^2}.\nonumber \end{align} In order to clearly show our work, we define three situations: cases A, B, and C. Cases B and C denote an asymmetric cavity with $\kappa_1=1.5\kappa_0$ and $\kappa_2=0.5\kappa_0$, corresponding to the forward input case ($a_{\rm{in}}$) and the backward input case ($b_{\rm{in}}$), respectively. Case A, independent of the input direction, denotes a symmetrical cavity with $\kappa_1=\kappa_2=\kappa_0$. Here, we give the specific parameter values used in this study, $\omega_{\rm m}/2\pi=20$ MHz, $\gamma_{\rm b}=100$ Hz, $g=300$ Hz, $\lambda_{\scriptscriptstyle{\rm L}}=2\pi c/\omega_{\scriptscriptstyle{\rm L}}=784.3$ nm, $J/2\pi=10$ MHz, and $\kappa_{\rm c}/2\pi=\kappa_{\rm a}/2\pi=2\kappa_0/2\pi=2$ MHz, which appeared in the previous research.[64,66,67] Note that $Q_{\rm c}$ is generally $10^5$–$10^{10}$, and $Q_{\rm m}$ is generally $10^5$–$10^6$.[16] Specifically, when the length of the cavity and the mass of the mechanical oscillator are $L=1$ mm and $m=10$ pg respectively, we can obtain that the optomechanical coupling strength is $g=\frac{\omega_{\rm c}}{L}\sqrt{\frac{\hbar}{2m\omega_{\rm m}}}\sim492$ Hz. In Fig. 2, we investigate the steady-state populations of intracavity photons $|c_{\rm s}|^2$ as a function of the optical detuning $\varDelta/\omega_{\rm m}$. It is obvious that the different driving directions (forward or backward) will lead to different steady-state photon numbers $|c_{\rm s}|^2$. The results show that compared with case A, the photon number of the asymmetric cavity in case B, $|c_{\rm s}|^2$, increases while it decreases in case C. This is due to the larger driving amplitude of case B compared to cases A and C. The radiation pressure is proportional to the intracavity photon numbers. Therefore, we can effectively adjust (increase or decrease) the intensity of optomechanical coupling by tuning the direction of the input light.
cpl-40-10-104201-fig2.png
Fig. 2. The steady-state photon number $|c_{\rm s}|^2$ as a function of the optical detuning $\varDelta/\omega_{\rm m}$. The dependence of photon number $|c_{\rm s}|^2$ on driving direction is obvious, i.e., the forward direction drive (case B) and the backward direction drive (case C) for $\kappa_1=1.5\kappa_0$ and $\kappa_2=0.5\kappa_0$. Case A ($\kappa_1=\kappa_2=\kappa_0$) is a symmetrical cavity. Here, the driving power is $P_{\rm{in}}=1\,µ$W. See text for the other parameters.
cpl-40-10-104201-fig3.png
Fig. 3. (a) The mechanical gain $G/\gamma_{\rm b}$ as a function of the optical detuning $\varDelta/\omega_{\rm m}$ in the different cases. The phonon laser can be generated (red dashed curve) or inhibited (green dotted curve) from different input directions. (b) The stimulated emitted phonon number $N_{\rm b}$ as a function of the pump power $P_{\rm{in}}$. The thick points correspond to the threshold power $P_{\rm{th}}$ at $\varDelta/\omega_{\rm m}=0.5$, which is determined by the threshold condition $G=\gamma_{\rm b}$. (c) Dependence of the isolation parameter $\Re$ on the decay rate $\kappa_1$. The other parameters are the same as those in Fig. 2.
We present the mechanical gain $G$ as a function of the optical detuning $\varDelta/\omega_{\rm m}$ for the different cases, as shown in Fig. 3(a). In case A (symmetrical cavity), the gain $G$ of the phonon laser is independent of the driving direction. It is a remarkable fact that only mechanical gain $G>\gamma_{\rm b}$ makes phonon lasing possible.[28] Since the driving strength is related to the driving direction in an asymmetric cavity, we can enhance the mechanical gain for case B (i.e., ${G}/\gamma_{\rm b}>1$) and apparently suppress it for case C (i.e., ${G}/\gamma_{\rm b} < 1$). Therefore, our scheme provides a novel way to control the phonon laser behavior. With the mechanical gain $G>\gamma_{\rm b}$ described above, the phonon mode can be amplified. Hence, we calculate the stimulated emitted phonon number $N_{\rm b}$, viz., \begin{align} N_{\rm b}=\exp[2(G-\gamma_{\rm b})/\gamma_{\rm b}]. \tag {10} \end{align} According to the threshold condition $G=\gamma_{\rm b}$, the corresponding threshold pump power $P_{\rm{th}}$ can be easily deduced as \begin{align} P_{\rm{th}}=C_i\cdot\dfrac{[4\kappa^2+(2J-\omega_{\rm m})^2](\beta^2+4\kappa^2\varDelta^2)}{4J\varDelta+2\varDelta(\omega_{\rm m}-2J)}, \tag {11} \end{align} where \begin{align} C_i=\dfrac{2\hbar\gamma_{\rm b}\omega_{\rm c}}{g^2\kappa\kappa_i},~~(i=0, 1, 2).\nonumber \end{align} The stimulated emitted phonon number $N_{\rm b}$ as a function of the driving power $P_{\rm{in}}$ with $J/2\pi=10$ MHz and $\varDelta/\omega_{\rm m}=0.5$ is shown in Fig. 3(b). Intuitively, the higher driving power can generate more phonons. In our scheme, the driving intensity is related to the loss rate of the cavity walls (i.e., $\rm{M_1}$ or $\rm{M_2}$) in Eq. (11). For the backward drive, the loss rate of the cavity wall is smaller, so a larger input power is needed to achieve a phonon laser. The power thresholds of cases A, B, and C are 0.72 µW, 0.48 µW, and 1.40 µW, respectively. For usable parameters, the power threshold of our results is smaller than the 7 µW reported in experiment and theory.[28,54]
cpl-40-10-104201-fig4.png
Fig. 4. The mechanical gain $G/\gamma_{\rm b}$ of the phonon laser as a function of the optical detuning $\varDelta/\omega_{\rm m}$ on (a) the atomic coupling strength $J$ and (b) the atomic loss rate $\kappa_{\rm a}$, respectively. The other parameters are the same as those in Fig. 2.
To clearly see the effect of our system on the stimulated emitted phonon, the relationship between $\kappa_1$ and $\kappa_2$ satisfies $(\kappa_1+\kappa_2)=2 \kappa_0$, where the average cavity-loss rate $\kappa_0/2\pi=1$ MHz is fixed. An isolation parameter can be introduced,[68] \begin{align} \Re=10\Big|\log_{10}\frac{N_{\rm b}(\kappa_1)}{N_{\rm b}(\kappa_2)}\Big|, \tag {12} \end{align} where nonzero $\Re$ indicates the occurrence of a nonreciprocal phonon laser. The isolation parameter $\Re$ as a function of loss rate of the cavity wall $\kappa_1$ for $\varDelta/\omega_{\rm m} = 0.5$ is shown in Fig. 3(c). For the phonon laser in a symmetrical cavity (i.e., case A for $\kappa_1=\kappa_2$), the isolation parameter is $\Re=0$ (the blue dot), which represents a reciprocal phonon laser. Next, we continue to discuss the asymmetrical cavity case (i.e., $\kappa_1\neq\kappa_2$), where $\Re\neq 0$ or $N_{\rm b}(\kappa_1) \neq N_{\rm b}(\kappa_2)$ indicates nonreciprocity driven from the two different directions. For convenience, we define $\kappa_1>\kappa_2$ as the forward input, i.e., case B ($1 < \kappa_1/\kappa_0 < 2$), and vice versa. Obviously, the phonon lasers for both cases (B and C) are dependent on the driving directions, which are indicated by the red dashed curve (for the forward direction) and the green dotted curve (for the backward direction) as shown in Fig. 3(c). The isolation degree of the nonreciprocal phonon laser can be adjusted by changing the loss ratio of the two cavity walls (i.e., $\rm{M_1}$ and $\rm{M_2}$), and the maximum isolation rate can reach approximately 24.2 dB. Our work makes it easy to turn the phonon laser on or off by adjusting the driving direction due to the asymmetric loss of cavity walls. Meanwhile, this result reveals one of the most important applications of nonreciprocal devices, as a means of protecting lasers from harmful reflections. Apart from that, we calculate the isolation parameter $\Re$ which varies linearly with the decay of the cavity wall (i.e., $\kappa_1$ or $\kappa_2$) in a robust way. This may provide a way to measure the loss ratio and quality factor of the cavity. Finally, we study the effect of atomic characteristics on the gain of the phonon laser in the symmetrical cavity (i.e., case A). In Fig. 4, we show the mechanical gain $G$ as a function of the optical detuning $\varDelta/\omega_{\rm m}$ for different values of the atomic coupling strength $J$ and the atomic loss rate $\kappa_{\rm a}$. We discuss the gain of the phonon laser in the case of $J/\omega_{\rm m}=0.45$, 0.50, and 0.55, as shown in Fig. 4(a), in which the atoms-photon coupling strength $J$ can be easily adjusted by the atomic number $N$. The mechanical gain decreases significantly except for the supermode resonance condition (i.e., $J/\omega_{\rm m}=0.5$), and the position of the maximum mechanical gain is $\varDelta=J$. In the experiment, the decay rate of the two-level atoms $\kappa_{\rm a}$ can be designed. A larger mechanical gain can be realized with smaller $\kappa_{\rm a}$ as shown in Fig. 4(b). It is explained that the smaller atomic dissipation leads to larger photon numbers, which increases the optomechanical coupling between the cavity field and mechanical mode. Importantly, it can be seen from Eq. (9) that the nonreciprocal phonon laser can still be realized, regardless of the atomic situation. This situation is not shown in the text. Our scheme provides a new way to regulate phonon lasers by adjusting the properties of atoms in the atoms-COM systems. In summary, we have theoretically studied a nonreciprocal phonon laser in a compound optomechanical system, consisting of an asymmetric COM resonator and an ensemble of $N$ two-level atoms. Compared with a symmetrical Fabry–Pérot cavity, the introduction of the asymmetric cavity walls (i.e., $\rm{M_1}$ and $\rm{M_2}$) significantly changes the steady-state photon number. The mechanical gain of the phonon laser in the asymmetric cavity system depends on the driving direction, making it possible to achieve the nonreciprocal phonon laser. The mechanical gain and the threshold power can be vigorously modulated by adjusting the loss rate $\kappa_1$ (or $\kappa_2$) of the cavity wall and the driving direction. Then, we also consider the effect of atomic properties on the mechanical gain. These results provide a novel path for manipulating the atoms-COM system through an asymmetric decay of the cavity walls, and may bring about interesting applications in various nonreciprocal devices. Acknowledgments. This work was supported by the National Key Research and Development Program of China (Grant No. 2021YFA1400700), the National Natural Science Foundation of China (Grant Nos. 11774113 and 12022507), and the Fundamental Research Funds for the Central Universities (Grant No. 2019kfyRCPY111).
References Cavity cooling of a microleverSub-kelvin optical cooling of a micromechanical resonatorBeating quantum limits in an optomechanical sensor by cavity detuningCavity optomechanical coupling assisted by an atomic gasQuantum Theory of Cavity-Assisted Sideband Cooling of Mechanical MotionResolved-sideband cooling of a micromechanical oscillatorSimultaneous cooling of coupled mechanical resonators in cavity optomechanicsPrecision measurement of electrical charge with optomechanically induced transparencyPrecision measurement of electrical charges in an optomechanical system beyond linearized dynamicsHighly sensitive optical sensor for precision measurement of electrical charges based on optomechanically induced difference-sideband generationOptically measuring force near the standard quantum limitDemonstration of Displacement Sensing of a mg-Scale Pendulum for mm- and mg-Scale Gravity MeasurementsPotential radiation-pressure-induced instabilities in cavity interferometersQuantum-Mechanical Radiation-Pressure Fluctuations in an InterferometerCavity Optomechanics: Back-Action at the MesoscaleCavity optomechanicsOptomechanically Induced TransparencyOptomechanically-induced transparency in parity-time-symmetric microresonatorsSpectrum and Dynamics of Optical Frequency Combs Generated with Monolithic Whispering Gallery Mode ResonatorsHigher-order sidebands in optomechanically induced transparencyPhoton Blockade Effect in Optomechanical SystemsNonreciprocal Photon BlockadeNonreciprocal photon blockade in a two-mode cavity with a second-order nonlinearityTunable photon statistics in parametrically amplified photonic moleculesWitnessing Optomechanical Entanglement with Photon CountingTwofold mechanical squeezing in a cavity optomechanical systemNonreciprocal microwave transmission under the joint mechanism of phase modulation and magnon Kerr nonlinearity effectPhonon Laser Action in a Tunable Two-Level System PT -Symmetric Phonon LaserDynamical phonon laser in coupled active-passive microresonatorsExceptional Points in Random-Defect Phonon LasersPhase-controlled phonon laserCatastrophic transition between dynamical patterns in a phonon laserA phonon laserOptically Driven Quantum Dots as Source of Coherent Cavity Phonons: A Proposal for a Phonon Laser SchemeProposal for a Phonon Laser Utilizing Quantum-Dot Spin StatesColloquium : Phononics: Manipulating heat flow with electronic analogs and beyondMode Competition and Anomalous Cooling in a Multimode Phonon LaserPhonon counting and intensity interferometry of a nanomechanical resonatorPhononic Frequency Comb via Intrinsic Three-Wave MixingQuantum optical circulator controlled by a single chirally coupled atomPhase-modulated single-photon nonreciprocal transport and directional router in a waveguide–cavity–emitter system beyond the chiral couplingInjection and detection of a spin-polarized current in a light-emitting diodeDesigning a Spin-Seebeck DiodeAll-passive nonreciprocal metastructureNonreciprocity Realized with Quantum NonlinearityApplications of magneto-optical waveguides in integrated optics: reviewObservation of Asymmetric Transport in Structures with Active NonlinearitiesParity–time-symmetric whispering-gallery microcavitiesMicroresonator isolators and circulators based on the intrinsic nonreciprocity of the Kerr effectMagnon-Induced Nonreciprocity Based on the Magnon Kerr EffectNon-reciprocal photonics based on time modulationNonreciprocal conventional photon blockade in driven dissipative atom-cavityNonreciprocal Phonon LaserNonreciprocal Optomechanical Entanglement against Backscattering LossesNonreciprocal chaos in a spinning optomechanical resonatorNonreciprocal sideband responses in a spinning microwave magnomechanical systemControllable optical bistability of an asymmetric cavity containing a single two-level atomRealization of Nonlinear Optical Nonreciprocity on a Few-Photon Level Based on Atoms Strongly Coupled to an Asymmetric CavityGiant nonreciprocal unconventional photon blockade with a single atom in an asymmetric cavityStrong dispersive coupling of a high-finesse cavity to a micromechanical membraneElectromagnetically induced absorption and transparency due to resonant two-field excitation of quasidegenerate levels in Rb vaporTunable slow and fast light in an atom-assisted optomechanical systemGeneration and amplification of a high-order sideband induced by two-level atoms in a hybrid optomechanical systemDicke quantum phase transition with a superfluid gas in an optical cavitySuperradiance-Driven Phonon LaserStatic non-reciprocity in mechanical metamaterials
[1] Metzger C H and Karrai K 2004 Nature 432 1002
[2] Kleckner D and Bouwmeester D 2006 Nature 444 75
[3] Arcizet O, Briant T, Heidmann A, and Pinard M 2006 Phys. Rev. A 73 033819
[4] Ian H, Gong Z R, Liu Y X, Sun C P, and Nori F 2008 Phys. Rev. A 78 013824
[5] Marquardt F, Chen J P, Clerk A A, and Girvin S M 2007 Phys. Rev. Lett. 99 093902
[6] Schliesser A, Rivière R, Anetsberger G, Arcizet O, and Kippenberg T J 2008 Nat. Phys. 4 415
[7] Lai D G, Zou F, Hou B P, Xiao Y F, and Liao J Q 2018 Phys. Rev. A 98 023860
[8] Zhang J Q, Li Y, Feng M, and Xu Y 2012 Phys. Rev. A 86 053806
[9] Xiong H, Si L G, and Wu Y 2017 Appl. Phys. Lett. 110 171102
[10] Xiong H, Liu Z X, and Wu Y 2017 Opt. Lett. 42 3630
[11] Schreppler S, Spethmann N, Brahms N, Botter T, Barrios M, and Stamper-Kurn D M 2014 Science 344 1486
[12] Matsumoto N, Lopez S B C N, Sugawara M, Suzuki S, Abe N, Komori K, Michimura Y, Aso Y, and Edamatsu K 2019 Phys. Rev. Lett. 122 071101
[13] Meers B and MacDonald N 1989 Phys. Rev. A 40 3754
[14] Caves C M 1980 Phys. Rev. Lett. 45 75
[15] Kippenberg T J and Vahala K J 2008 Science 321 1172
[16] Aspelmeyer M, Kippenberg T J, and Marquardt F 2014 Rev. Mod. Phys. 86 1391
[17] Weis S, Rivière R, Deléglise S, Gavartin E, Arcizet O, Schliesser A, and Kippenberg T J 2010 Science 330 1520
[18] Jing H, Özdemir Ş K, Geng Z, Zhang J, Lü X Y, Peng B, Yang L, and Nori F 2015 Sci. Rep. 5 9663
[19] Chembo Y K, Strekalov D V, and Yu N 2010 Phys. Rev. Lett. 104 103902
[20] Xiong H, Si L G, Zheng A S, Yang X, and Wu Y 2012 Phys. Rev. A 86 013815
[21] Rabl P 2011 Phys. Rev. Lett. 107 063601
[22] Huang R, Miranowicz A, Liao J Q, Nori F, and Jing H 2018 Phys. Rev. Lett. 121 153601
[23] Wang K, Wu Q, Yu Y F, and Zhang Z M 2019 Phys. Rev. A 100 053832
[24] Shen S T, Qu Y, Li J H, and Wu Y 2019 Phys. Rev. A 100 023814
[25] Ho M, Oudot E, Bancal J D, and Sangouard N 2018 Phys. Rev. Lett. 121 023602
[26] Hu C S, Yang Z B, Wu H, Li Y, and Zheng S B 2018 Phys. Rev. A 98 023807
[27] Kong C, Liu J, and Xiong H 2023 Front. Phys. 18 12501
[28] Grudinin I S, Lee H, Painter O, and Vahala K J 2010 Phys. Rev. Lett. 104 083901
[29] Jing H, Özdemir S K, Lü X Y, Zhang J, Yang L, and Nori F 2014 Phys. Rev. Lett. 113 053604
[30] He B, Yang L, and Xiao M 2016 Phys. Rev. A 94 031802
[31] Lü H, Özdemir S K, Kuang L M, Nori F, and Jing H 2017 Phys. Rev. Appl. 8 044020
[32] Zhang Y L, Zou C L, Yang C S, Jing H, Dong C H, Guo G C, and Zou X B 2018 New J. Phys. 20 093005
[33] Lin Q, He B, and Xiao M 2021 Phys. Rev. Res. 3 L032018
[34] Vahala K, Herrmann M, Knünz S, Batteiger V, Saathoff G, Hänsch T W, and Udem T 2009 Nat. Phys. 5 682
[35] Kabuss J, Carmele A, Brandes T, and Knorr A 2012 Phys. Rev. Lett. 109 054301
[36] Khaetskii A, Golovach V N, Hu X, and Žutić I 2013 Phys. Rev. Lett. 111 186601
[37] Li N B, Ren J, Wang L, Zhang G, Hänggi P, and Li B W 2012 Rev. Mod. Phys. 84 1045
[38] Kemiktarak U, Durand M, Metcalfe M, and Lawall J 2014 Phys. Rev. Lett. 113 030802
[39] Cohen J D, Meenehan S M, MacCabe G S, Gröblacher S, Safavi-Naeini A H, Marsili F, Shaw M D, and Painter O 2015 Nature 520 522
[40] Ganesan A, Do C, and Seshia A 2017 Phys. Rev. Lett. 118 033903
[41] Scheucher M, Hilico A, Will E, Volz J, and Rauschenbeutel A 2016 Science 354 1577
[42] Wang X, Yang W X, Chen A X, Li L, Shui T, Li X, and Wu Z 2022 Quantum Sci. Technol. 7 015025
[43] Fiederling R, Keim M, Reuscher G, Ossau W, Schmidt G, Waag A, and Molenkamp L W 1999 Nature 402 787
[44] Borlenghi S, Wang W, Fangohr H, Bergqvist L, and Delin A 2014 Phys. Rev. Lett. 112 047203
[45] Mahmoud A M, Davoyan A R, and Engheta N 2015 Nat. Commun. 6 8359
[46] Hamann A R, Müller C, Jerger M, Zanner M, Combes J, Pletyukhov M, Weides M, Stace T M, and Fedorov A 2018 Phys. Rev. Lett. 121 123601
[47] Dötsch H, Bahlmann N, Zhuromskyy O, Hammer M, Wilkens L, Gerhardt R, Hertel P, and Popkov A F 2005 J. Opt. Soc. Am. B 22 240
[48] Bender N, Factor S, Bodyfelt J D, Ramezani H, Christodoulides D N, Ellis F M, and Kottos T 2013 Phys. Rev. Lett. 110 234101
[49] Peng B, Özdemir K, Lei F, Monifi F, Gianfreda M, Long G L, Fan S, Nori F, Bender C M, and Yang L 2014 Nat. Phys. 10 394
[50] Del Bino L, Silver J M, Woodley M T M, Stebbings S L, Zhao X, and Del'Haye P 2018 Optica 5 279
[51] Kong C, Xiong H, and Wu Y 2019 Phys. Rev. Appl. 12 034001
[52] Sounas D L and Alú A 2017 Nat. Photon. 11 774
[53] Xue W S, Shen H Z, and Yi X X 2020 Opt. Lett. 45 4424
[54] Jiang Y, Maayani S, Carmon T, Nori F, and Jing H 2018 Phys. Rev. Appl. 10 064037
[55] Jiao Y F, Zhang S D, Zhang Y L, Miranowicz A, Kuang L M, and Jing H 2020 Phys. Rev. Lett. 125 143605
[56] Zhang D W, Zheng L L, You C, Hu C S, Wu Y, and Lü X Y 2021 Phys. Rev. A 104 033522
[57] Wang X, Huang K W, and Xiong H 2023 Opt. Express 31 5492
[58] Xia X W, Xu J P, and Yang Y P 2014 Phys. Rev. A 90 043857
[59] Yang P F, Xia X W, He H, Li S K, Han X, Zhang P, Li G, Zhang P F, Xu J P, Yang Y P, and Zhang T C 2019 Phys. Rev. Lett. 123 233604
[60] Xia X W, Zhang X Q, Xu J P, Li H Z, Fu Z Y, and Yang Y P 2021 Phys. Rev. A 104 063713
[61] Thompson J D, Zwickl B M, Jayich A M, Marquardt F, Girvin S M, and Harris J G E 2008 Nature 452 72
[62] Akulshin A M, Barreiro S, and Lezama A 1998 Phys. Rev. A 57 2996
[63] Gu K H, Yan X B, Zhang Y, Fu C B, Liu Y M, Wang X, and Wu J H 2015 Opt. Commun. 338 569
[64] Liu Z X, Xiong H, and Wu Y 2018 Phys. Rev. A 97 013801
[65]Gardiner C, Zoller P, and Zoller P 2004 Quantum Noise: A Handbook of Markovian and Non-Markovian Quantum Stochastic Methods with Applications to Quantum Optics (Berlin: Springer Science & Business Media)
[66] Baumann K, Guerlin C, Brennecke F, and Esslinger T 2010 Nature 464 1301
[67] Jiang Y J, Lv H, and Jing H 2018 Chin. Phys. Lett. 35 044205
[68] Coulais C, Sounas D, and Alù A 2017 Nature 542 461