• Photonics Research
  • Vol. 8, Issue 2, 143 (2020)
Xunwei Xu1、*, Yanjun Zhao2, Hui Wang3, Hui Jing4、6, and Aixi Chen1、5、7
Author Affiliations
  • 1Department of Applied Physics, East China Jiaotong University, Nanchang 330013, China
  • 2Faculty of Information Technology, College of Microelectronics, Beijing University of Technology, Beijing 100124, China
  • 3Center for Emergent Matter Science (CEMS), RIKEN, Wako, Saitama 351-0198, Japan
  • 4Key Laboratory of Low-Dimensional Quantum Structures and Quantum Control of Ministry of Education, Department of Physics and Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha 410081, China
  • 5Department of Physics, Zhejiang Sci-Tech University, Hangzhou 310018, China
  • 6e-mail: jinghui73@foxmail.com
  • 7e-mail: aixichen@zstu.edu.cn
  • show less
    DOI: 10.1364/PRJ.8.000143 Cite this Article Set citation alerts
    Xunwei Xu, Yanjun Zhao, Hui Wang, Hui Jing, Aixi Chen. Quantum nonreciprocality in quadratic optomechanics[J]. Photonics Research, 2020, 8(2): 143 Copy Citation Text show less

    Abstract

    We propose to achieve nonreciprocal quantum control of photons in a quadratic optomechanical (QOM) system based on directional nonlinear interactions. We show that by optically pumping the QOM system in one side, the effective QOM coupling can be enhanced significantly in that side, but not for the other side. This, contrary to the intuitive picture, allows the emergence of a nonreciprocal photon blockade in such optomechanical devices with weak single-photon QOM coupling. Our proposal opens up the prospect of exploring and utilizing quantum nonreciprocal optomechanics, with applications ranging from single-photon nonreciprocal devices to on-chip chiral quantum engineering.

    1. INTRODUCTION

    Optomechanical systems can be used to test the foundations of physics, and have important applications ranging from gravitational wave detections to quantum information processing (for reviews, see Refs. [16]). Recently, optomechanically induced nonreciprocity has been proposed theoretically [718] and demonstrated experimentally [1928]. Nonreciprocal devices [29], such as isolators and circulators, have drawn an immense amount of interest in the past few years, for their irreplaceable role in signal processing and communication. An optomechanical system opens a route for nonreciprocity with nonlinear interactions [7,8,19,20], inherent non-trivial topology [9,2124], and synthetic magnetism [1018,2528].

    In this paper, we show that nonreciprocity can be realized in a whispering-gallery-mode (WGM) optomechanical system with quadratic optomechanical (QOM) coupling for quadratic mechanical-position dependent change in the optical frequency. Specifically, by optically pumping the quadratic WGM optomechanical system in one side, the effective QOM coupling can be enhanced significantly in that side, but not for the other side. The directional nonlinear interactions can induce nonreciprocal photon transport. This new possibility, as far as we know, has not been revealed in previous works.

    It has been shown that the QOM systems driven by a strong optical driving field can generate strong (second-order) nonlinear photon-phonon interaction, even under weak single-photon optomechanical coupling conditions [30,31]. Nevertheless, we show here that the WGM optomechanical system with QOM coupling can be used to generate directional nonlinear interactions, which opens up the prospect of exploring nonreciprocal photon transport based on nonlinearity without the limitations of dynamic reciprocity [32]. Moreover, we demonstrate that directional nonlinear interactions can not only induce classical nonreciprocity, but also achieve nonreciprocal quantum control of photons by manipulating the statistic of the nonreciprocal transport photons.

    Recently, a quantum nonreciprocal effect called nonreciprocal photon blockade (PB) was predicted theoretically [33] in a spinning Kerr resonator [3437], that PB can emerge when the resonator is driven in one direction but not the other. Different from the nonreciprocal PB induced by the Fizeau–Sagnac drag [33], which relies on the condition of strong single-photon nonlinearity, here we show that the nonreciprocal PB can also be explored even with weak single-photon QOM coupling. Our proposal provides a feasible method to realize nonreciprocal photon blockade.

    We note that the WGM optomechanical system has already been used for nonreciprocity with inherent non-trivial topology [9,2124]. However, the WGM optomechanical systems with linear mechanical-position dependent change in the optical frequency (i.e., linear optomechanical coupling) can only enhance linear photon-phonon interaction, which cannot be used to observe nonreciprocal PB. Our work provides not only a new platform (quadratic WGM optomechanics) but a new approach (i.e., directional nonlinear interaction) to designing nonreciprocal single-photon devices for backaction-immune quantum communication and chiral quantum photonics.

    2. MODEL AND HAMILTONIAN

    We note that the cavity frequency of a WGM optomechanical system is almost linearly proportional to the mechanical position [38,39], even though the effects of QOM coupling have also been observed in experiments [40,41]. The QOM coupling can facilitate nondestructive measurements of the energy of the mechanical oscillator [42,43], which is instead impossible for a system with linear optomechanical coupling. We present a scheme to generate QOM interactions in the normal optical modes of a WGM optomechanical system by eliminating the linear optomechanical couplings [4447].

    We consider a near-field cavity optomechanical setup consisting of one mechanical resonator optomechanically coupling to two optical resonators (j=1,2) via the optical evanescent field, with each optical resonator supporting two degenerate clockwise (CW) and counter-clockwise (CCW) travelling-wave WGMs, as shown in Figs. 1(a) and 1(b). The model can be described by the optomechanical interaction Hamiltonian Hom=j=1,2λ=cw,ccw[ω0+(1)jg0q]aj,λaj,λ+J(a1,ccwa2,cw+a1,cwa2,ccw+H.c.)+12ωm(q2+p2),where aj,λ and aj,λ (j=1,2 and λ=cw,ccw) are the annihilation and creation operators of the optical modes with frequency ω0, and q and p are the dimensionless displacement and momentum operators of the mechanical resonator with frequency ωm. As shown in Figs. 1(a) and 1(a), the mechanical resonator is placed at the middle of the gap of the two WGM resonators, so that the mechanical mode couples to the two WGM optical modes with linear optomechanical coupling of equal magnitude g0 but opposite signs. Due to the tunneling coupling J between the optical modes, there is an avoided crossing for the optical resonance frequency, ω±(q)=ω0±J2+(g0q)2, as shown in Figs. 1(c) and 1(d). We assume that the optical mode aj,λ is coupled to a waveguide (port j) with strength γc, and the damping rate of mechanical resonator q is γm. This theoretical model may also be realized with a membrane suspended on WGM microcavities [48] or photonic-crystal microcavities [49].

    (a), (b) Schematic diagram for generating QOM coupling, where a mechanical nanostring oscillator is placed between two whispering gallery mode (WGM) resonators. (c), (d) Dispersion of the optical modes as a function of the displacement.

    Figure 1.(a), (b) Schematic diagram for generating QOM coupling, where a mechanical nanostring oscillator is placed between two whispering gallery mode (WGM) resonators. (c), (d) Dispersion of the optical modes as a function of the displacement.

    Following the approach in Refs. [4447] (see Appendix A for more details), where |J|ωm is assumed such that q can be treated as a quasi-static variable, the Hamiltonian can be diagonalized quasi-statically as Hom(ω++gq2)(aL,+aL,++aR,+aR,+)+(ωgq2)(aL,aL,+aR,aR,)+12ωm(q2+p2)for normal modes aR,±(a1,cw±a2,ccw)/2 and aL,±(a1,ccw±a2,cw)/2 with eigenfrequencies ω±(q)ω0±(J+gq2) and single-photon QOM coupling gg02/(2J).

    As already shown in the experiment [47], when the tunneling coupling between the optical modes is strong, i.e., J>γc, the transmission spectrum of a laser probe through the coupled optical modes features resonance dips at the normal resonance frequencies ω±=ω0±J, not at the bare optical resonance frequencies ω0. That is to say, the normal modes are coupled to the external waveguides and can be used to describe the input–output characteristic of the coupled optical modes system. Specifically, the normal modes aL,± and aR,± are coupled to both of the ports with strength γc/2, respectively, and the total damping rates are γc.

    We can choose either pair of degenerated quasi-static normal optical modes (aL,+ and aR,+, or aL, and aR,), to generate strong nonlinear interaction for a few photons travelling in one direction, but not in the reverse direction. Without loss of generality, the two optical modes are denoted as aL and aR, with frequency ωa=ω+ or ω, for photons travelling from port 1 to port 2 and the opposite direction, respectively. For example, to enhance the nonlinear optomechanical coupling between the optical mode aL and the mechanical resonator b, the optical mode aL is pumped by a strong field with amplitude Ω and frequency ωLωa2ωm. In the rotating reference frame with the optical frequency ωL, the system can be described by Hom+driv=(Δa+gq2)(aLaL+aRaR)+12ωm(q2+p2)+ΩaL+ΩaL,with detuning Δa=ωaωL. Under the strong driving condition Ωγc, we perform the displacement transformations: aLαL+aL, aRαR+aR, qqs+q, and pps+p, where αL, αR, qs, and ps are the steady state values, and aL, aR, q, and p (on the right side of the arrow) are the quantum fluctuation operators. Adopting the standard linearization technique (see Appendix B), an effective Hamiltonian [30,31] for the quantum fluctuation operators is obtained as Heff=Δa(aLaL+aRaR)+ωmbb+GaLb2+G*aLb2,where ωm=ωm+g|αL|2 is the effective mechanical frequency, and G=gαL/2 is the effective (second-order) nonlinear coupling strength between the optical and mechanical modes. Without loss of generality, G is assumed to be real in the following.

    To investigate the system’s response behavior to a weak probe field, a weak field with amplitude εγc and frequency ωpωa is input from one of the ports. In the rotating reference frame with the unitary operator R(t)=exp[iδ(aLaL+aRaR+bb/2)t], the total Hamiltonian is given by Htot=Heff+Hprobe, where Heff=ΔaLaL+ΔaRaR+Δmbb+GaLb2+G*aLb2,with detunings δ=ωpωL, Δ=Δaδ, and Δm=ωm+g|αL|2δ/2 satisfying the condition max{|Δ|,|Δm|}ωm. Hprobe=εaξ+H.c. describes the driving of the probe field with subscript ξ=L for weak field input from port 1 and ξ=R for weak field input from port 2. The effective Hamiltonian Heff describes a directional nonlinear interaction between the optical and mechanical modes. If a probe field is input from port 1, the enhanced nonlinear interaction G between aL and b may suppress the excitation of the second photon under a single-photon resonant condition. In contrast, by driving from port 2, as there is no strong nonlinear interaction between aR and b, the photons are free to cross the system.

    According to the input–output relations [50], we have a1,in=ε/γc/2 and a2,out=γc/2aL (a2,in=ε/γc/2 and a1,out=γc/2aR), and then the transmission coefficient for the weak probe field can be defined by T21a2,outa2,outa1,ina1,in=γc24ε2aLaL,for photon transport from port 1 to port 2, and T12a1,outa1,outa2,ina2,in=γc24ε2aRaR,for photon transport from port 2 to port 1, where nLaLaL and nRaRaR are the mean photon numbers. The isolation for probe field transport from port 1 to port 2 is defined by IT21/T12. Using the input–output relations, the statistic properties of the transmitted photons a2,out and a1,out can be described by the second-order correlation functions in the steady state (t) as follows: g21(2)(τ)a2,out(t)a2,out(t+τ)a2,out(t+τ)a2,out(t)a2,out(t)a2,out(t)2=aL(t)aL(t+τ)aL(t+τ)aL(t)aL(t)aL(t)2,for photon transport from port 1 to port 2, and g12(2)(τ)a1,out(t)a1,out(t+τ)a1,out(t+τ)a1,out(t)a1,out(t)a1,out(t)2=aR(t)aR(t+τ)aR(t+τ)aR(t)aR(t)aR(t)2,for photon transport from port 2 to port 1.

    3. NONRECIPROCAL PB

    The transmission coefficients and correlation functions can be obtained by numerically solving the master equation for the density matrix ρ of the system [51] ρt=i[Htot,ρ]+γcL[aL]ρ+γcL[aR]ρ+γm(nth+1)L[b]ρ+γmnthL[b]ρ,where L[o]ρ=oρo(ooρ+ρoo)/2 denotes a Lindbland term for an operator o, and nth is the mean thermal phonon number, given by the Bose–Einstein statistics nth=[exp(ωm/kBT)1]1 with the Boltzmann constant kB and the temperature T of the reservoir at the thermal equilibrium.

    In Fig. 2(a), we show the transmission coefficients T21 for probe field transport from port 1 to port 2, and T12 for probe field transport from port 2 to port 1 versus the detuning Δ/G. For T21, there are two peaks at Δ=±2G and one dip at Δ=0. In contrast, there is only one peak at Δ=0 for T12. Figure 2(b) shows the isolation I=T21/T12 for probe field transport from port 1 to port 2 as a function of the detuning Δ/G. The isolation for the direction 12 is more than 17 dB at detuning Δ=±2G, and the isolation for the reverse direction 21, i.e., I˜T12/T21, is more than 70 dB at detuning Δ=0.

    (a) The transmission coefficients T21 (solid black curve) and T12 (dashed red curve) as a function of the detuning Δ/G. (b) The isolation as a function of the detuning Δ/G. (c) The equal-time second-order correlation function log10[gij(2)(0)] (ij=12,21) as a function of the detuning Δ/G. (d) The second-order correlation function log10[g21(2)(τ)] as a function of the normalized time delay γcτ/(2π) at detuning Δ=2G. The other parameters are Δm=Δ/2, G=3γc, ε=γc/10, γm=γc/100, and nth=0.

    Figure 2.(a) The transmission coefficients T21 (solid black curve) and T12 (dashed red curve) as a function of the detuning Δ/G. (b) The isolation as a function of the detuning Δ/G. (c) The equal-time second-order correlation function log10[gij(2)(0)] (ij=12,21) as a function of the detuning Δ/G. (d) The second-order correlation function log10[g21(2)(τ)] as a function of the normalized time delay γcτ/(2π) at detuning Δ=2G. The other parameters are Δm=Δ/2, G=3γc, ε=γc/10, γm=γc/100, and nth=0.

    To explore the statistic properties of the transmitted photons, the equal-time second-order correlation function log10[gij(2)(0)] (ij=12,21) is shown as a function of the detuning Δ/G in Fig. 2(c). The photon transport from port 2 to port 1 is coherent in full frequency, i.e., g12(2)(0)=1. The photon transport from port 1 to port 2 exhibits strong antibunching effect, i.e., g21(2)(0)1, around the detunings Δ=±2G, and exhibits strong bunching effect, i.e., g21(2)(0)1, around the detunings Δ=0 and ±2G. The time duration for nonreciprocal PB at Δ=±2G is on the order of 2π/(10γc), as shown in Fig. 2(d).

    The peak for T121 at detuning Δ=0 can be understood by the fact that when the probe field is injected from port 2, only the (linear) optical mode aR can be excited. Thus the maximum transmission coefficient is reached for the probe field in resonance with the optical mode aR, i.e., Δ=0, and the transmitted photons keep the statistic properties of the probe field (a coherent field), i.e., g12(2)(0)=1, as there is no nonlinear interaction in the optical path from port 2 to port 1.

    In order to understand the origin of the peak for T210.8 around the detuning Δ=±2G and the dip for T21107 at Δ=0, we show energy spectrum (see Appendix C) of the linearized QOM coupling between optical mode aL and mechanical resonator b in Fig. 3. Under weak probe condition, the maximum transmission coefficient T210.8 is reached for the probe field in resonance with the transition |00|2±1, i.e., Δ=±2G, as shown in Fig. 3(a). However, the photon absorbed in the transition |00|2±1 blocks the transition |2±1|4±1 for large detuning, so we have g12(2)(0)1 around the detuning Δ=±2G. The dip for T21107 at Δ=0 arises from the quantum interference between the transitions |2+1|00 and |21|00 [Fig. 3(b)], in an equivalent picture as optomechanically induced transparency [5254], or electromagnetically induced transparency in lambda-type three-level atoms [55,56]. Moreover, when Δ=0, the transition |00|2±1 is suppressed, but the two-photon transition |00|40 is resonant, which induces two-photon tunneling from port 1 to point 2, i.e., g21(2)(0)1. Similarly, g21(2)(0)1 around Δ=±2G is induced by the resonant transition |00|4±.

    Schematic energy spectrum of the linearized QOM coupling between optical mode aL and mechanical resonator b, where |00⟩≡|0,0⟩, |10⟩≡|0,1⟩, |2±1⟩≡(|1,0⟩±|0,2⟩)/2, |3±1⟩≡(|1,1⟩±|0,3⟩)/2, |40⟩≡(−3|2,0⟩+|0,4⟩)/2, |4±1⟩≡(|2,0⟩±2|1,2⟩+3|0,4⟩)/(22), and |n,m⟩ represents the Fock state with n photons in aL and m phonons in b.

    Figure 3.Schematic energy spectrum of the linearized QOM coupling between optical mode aL and mechanical resonator b, where |00|0,0, |10|0,1, |2±1(|1,0±|0,2)/2, |3±1(|1,1±|0,3)/2, |40(3|2,0+|0,4)/2, |4±1(|2,0±2|1,2+3|0,4)/(22), and |n,m represents the Fock state with n photons in aL and m phonons in b.

    We now discuss the effect of thermal phonons on the nonreciprocal PB. Figures 4(a) and 4(b) show the transmission coefficient T21 and the equal-time second-order correlation function log10[g21(2)(0)] versus the detuning Δ/G with a different mean thermal phonon number (nth=0,0.1,1). Thermal phonons have little influence on the transmission coefficient T21 around Δ=±2G, but have great effect on the transmission coefficient T21 around Δ=0. Thermal phonons have great effect on the second-order correlation function log10[g21(2)(0)] around Δ=0 and ±2G.

    (a) Transmission coefficient T21. (b) The equal-time second-order correlation function log10[g21(2)(0)] versus the detuning Δ/G with different mean thermal phonon number (nth=0,0.1,1). (c) The isolation T21/T12. (d) The equal-time second-order correlation function log10[g21(2)(0)] versus the mean thermal phonon number nth with different detuning (Δ=0,2G,6G). The other parameters are the same as in Fig. 2.

    Figure 4.(a) Transmission coefficient T21. (b) The equal-time second-order correlation function log10[g21(2)(0)] versus the detuning Δ/G with different mean thermal phonon number (nth=0,0.1,1). (c) The isolation T21/T12. (d) The equal-time second-order correlation function log10[g21(2)(0)] versus the mean thermal phonon number nth with different detuning (Δ=0,2G,6G). The other parameters are the same as in Fig. 2.

    The isolation T21/T12 and the second-order correlation function log10[g21(2)(0)] versus the mean thermal phonon number nth are shown in Figs. 4(c) and 4(d) with detuning Δ=0,2G,6G. The isolation T21/T12 around Δ=2G is robust against the thermal phonons, but the antibunching effect of the transport photons becomes much weaker for greater thermal phonons. Both the isolation T21/T12 and second-order correlation function log10[g21(2)(0)] around Δ=0 are sensitive to the mean thermal phonon number nth, and this quality may be used in accurate temperature measurement at ultra-low temperature.

    More interestingly, two peaks appear around Δ=±6G in the transmission coefficient T21, and the isolation T21/T12 can be improved with a larger thermal phonon number nth. This abnormal effect is induced by the phonon states, e.g., |10 in Fig. 3(c). As the temperature increases, the population probability in phonon state |10 increases, and the transitions of |10|3±1 with resonance frequency Δ=±6G become remarkable gradually, which induces the increasing peaks of the transmission coefficient T21 (or the isolation T21/T12) around Δ=±6G.

    4. DISCUSSIONS AND CONCLUSIONS

    Let us now discuss the experimental requirements for our proposal. Considering the parameters from a recent near-field WGM optomechanical experiment [39]: g02π×(10100)  kHz, γc2π×(1001000)  MHz, and J2π×1000  MHz for strong coupling, we have QOM coupling g/2π10  Hz, which is too weak to realize G=gαL/2>γc. However, many schemes have been proposed to enhance the optomechanical coupling g0 in recent years, such as by using the collective mechanical modes in scatter arrays [57], the squeezed cavity mode [58], or the Josephson effect for electromechanical systems [59]. The optomechanical coupling g0 can be enhanced by several orders of magnitude and the strong coupling regime (g0>γc) can even be approached [5759]. Furthermore, the frequency range of the SiN nanostring oscillators in the near-field cavity optomechanical setup is about 0.1–10 MHz [3841], which is too low to observe nonreciprocal PB. One possible solution is replaying the SiN nanostring oscillator with an optomechanical crystal nanobeam, as optomechanical crystal nanobeams with GHz mechanical modes have been designed and demonstrated recently [60,61]. Moreover, the thermal phonons have important influence on the nonreciprocal PB, and the thermal effect can be suppressed for higher frequency mechanical mode.

    In addition, there are two driving fields coupling to the system with frequencies ωLωa2ωm and ωpωa, and the strengths Ωε. In order to detect the photon correlation induced by the weak probe field, we should spectrally filter out the strong optical driving field at frequency ωL under the condition ωaωL{γc,G}. A similar method has already been used to measure correlations of phonons in a recent experiment [62].

    In summary, we have shown that the WGM optomechanical system with QOM coupling can be used to obtain directional nonlinear interactions and observe nonreciprocal PB. In addition, the thermal phonons have important influence on the nonreciprocal PB, especially on the statistic properties of the transport photons, and this quality may be used in temperature sensing at ultra-low temperature. Moreover, this work can be extended to study phonon manipulation in double-cavity optomechanics, e.g., nonreciprocal phonon blockades [63], nonreciprocal phonon lasers [6470], nonreciprocal photon-phonon entanglements and quantum transfers. Directional nonlinear interaction can also be applied in the high-Q microtoroid resonator with phase-matched parametric amplification [71] or superconducting microwave circuit with Josephson parametric converters [72] to achieve nonreciprocal quantum control of photons. Our proposal provides a new routine towards the realization of on-chip quantum nonreciprocal devices, backaction-immune quantum measurement, and chiral quantum physics.

    APPENDIX A: QUADRATIC OPTOMECHANICAL COUPLING

    To realize quantum nonreciprocality in optomechanics, we consider a near-field cavity optomechanical setup consisting of one mechanical resonator optomechanical coupling to two optical resonators (j=1,2) via the optical evanescent field, with each optical resonator supporting two degenerate clockwise (CW) and counter-clockwise (CCW) travelling-wave whispering-galley-modes (WGMs). The model can be described by the optomechanical interaction Hamiltonian Hom=(a1,ccw a2,cw )(ω0 g0qJJω0+g0q)(a1,ccwa2,cw)+(a1,cw a2,ccw )(ω0 g0qJJω0+g0q)(a1,cwa2,ccw)+12ωm(q2+p2),where aj,λ and aj,λ (j=1,2 and λ=cw,ccw) are the annihilation and creation operators of the optical modes with frequency ω0, J is the tunneling amplitude between the optical modes, q and p are the dimensionless displacement and momentum operators of the mechanical resonator with frequency ωm, and g0 is the linear optomechanical coupling strength between the mechanical resonator and optical modes. We assume that the optical mode aj,λ is coupled to a waveguide (port j) with strength γc, and the damping rate of mechanical resonator q is γm.

    The Hamiltonian in Eq. (A1) can be diagonalized as Hom=(aL,+ aL, )(ω+(q)00ω (q))(aL,+aL, )+(aR,+ aR, )(ω+(q)00ω (q))(aR,+aR, )+12ωm(q2+p2),in the normal modes basis, aL,±=1D±[Ja1,ccw+(g0q±J2+(g0q)2)a2,cw]and aR,±=1D±[Ja1,cw+(g0q±J2+(g0q)2)a2,ccw],with D±2=J2+(g0q±J2+(g0q)2)2,and eigenfrequencies ω±(q)=ω0±J2+(g0q)2.

    Moreover, |J| g0q is assumed [4447] such that we can Taylor-expand the eigenfrequencies in Eq. (A6) as ω±(q)ω±±g022Jq2,with frequencies ω±ω±(0)=ω0±J, and the quasi-static Hamiltonian in Eq. (A2) is given approximately by Eq. (2) in the main text with quadratic optomechanical coupling gg02/(2J), and quasi-static normal optical modes, aL,±(a1,ccw±a2,cw)/2 and aR,±(a1,cw±a2,ccw)/2.

    APPENDIX B: LINEARIZED OPTOMECHANICAL HAMILTONIAN

    We consider a pair of degenerated quasi-static normal optical modes (aL,+ and aR,+, or aL, and aR, ), to generate strong nonlinear interaction for a few photons travelling in one direction, but not in the reverse direction. Without loss of generality, the two optical modes are denoted as aL and aR, with frequency ωa=ω+ or ω , for photons travelling from port 1 to port 2 and the opposite direction, respectively. For example, to enhance the nonlinear optomechanical coupling between the optical mode aL and the mechanical resonator, the optical mode aL is pumped by a strong field with amplitude Ω and frequency ωLωa 2ωm. In the rotating reference frame with the optical frequency ωL, the system can be described by Eq. (3) in the main text with detuning Δa=ωa ωL.

    Under strong driving condition Ω γc, we perform the displacement transformations: aLαL+aL, aRαR+aR, qqs+q, and pps+p, where αL, αR, qs, and ps are the steady state values, and aL, aR, q, and p (on the right side of the arrow) are the quantum fluctuation operators. The steady state values αL, αR, qs, and ps can be obtained by the equations of motions yielding αL= i2Ωγc+i2Δaand αR=qs=ps=0.

    The operators q and p for the mechanical resonator can be written in terms of phonon creation and annihilation operators as q=(b +b)/2, p=i(b b)/2, and the effective Hamiltonian for the quantum fluctuation operators reads Heff=ΔaaL aL+ΔaaR aR+ωmb b+g2(|αL|2+aL aL+aR aR)(b +b)2+g2(αLaL +αL*aL)(b +b)2.

    We assume that the optical driving field is strong enough so we have |αL|2 aL aL aR aR , and so that the term g(aL aL+aR aR)(b +b)2 in the above equation can be neglected safely. For Δa2ωm |gαL|/2, the effective Hamiltonian can be further simplified by rotating-wave approximation and neglecting the high frequency terms [30,31], e.g., b2 and ab2. Then, we get Eq. (4) in the main text, where ωm=ωm+g|αL|2 is the effective mechanical frequency, and G=gαL/2 is the effective nonlinear coupling strength between the optical and mechanical modes.

    APPENDIX C: ENERGY SPECTRUM

    In this appendix, we will show the energy spectrum of the effective Hamiltonian in the diagonal basis under the condition Δ=2Δm, and similar results are shown in Refs. [30,31]. In the noncoupling basis, |n,m represents the Fock state with n photons in aL and m phonons in b. It is clear that the vacuum state |00 |0,0 and the single-phonon state |10 |0,1 ,are decoupled from the other states. The degenerate states |1,0 and |0,2 are coupled with strengh 2G, so the eigenstates are |2±1 12(|1,0 ±|0,2 ),with energy splitting Δ2,±=±2G.

    The degenerate states |1,1 and |0,3 are coupled with strength 6G, so the eigenstates are |3±1 12(|1,1 ±|0,3 ),with energy splitting Δ3,±=±6G.

    The states |2,0 , |1,2 , and |0,4 are degenerate, where |2,0 and |1,2 are coupled with strength 2G, and |0,4 and |1,2 are coupled with strength 23G, so the eigenstates are |40 12( 3|2,0 +|0,4 ),|4±1 122(|2,0 ±2|1,2 +3|0,4 ),with energy splitting Δ4,0=0,Δ4,±=±4G.

    The schematic energy spectrum in the diagonal basis is given in Fig. 3.

    References

    [1] T. J. Kippenberg, K. J. Vahala. Cavity optomechanics: back-action at the mesoscale. Science, 321, 1172-1176(2008).

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

    [3] M. Aspelmeyer, P. Meystre, K. Schwab. Quantum optomechanics. Phys. Today, 65, 29-35(2012).

    [4] M. Aspelmeyer, T. J. Kippenberg, F. Marquardt. Cavity optomechanics. Rev. Mod. Phys., 86, 1391-1452(2014).

    [5] M. Metcalfe. Applications of cavity optomechanics. Appl. Phys. Rev., 1, 031105(2014).

    [6] Y.-L. Liu, C. Wang, J. Zhang, Y.-X. Liu. Cavity optomechanics: manipulating photons and phonons towards the single-photon strong coupling. Chin. Phys. B, 27, 024204(2018).

    [7] S. Manipatruni, J. T. Robinson, M. Lipson. Optical nonreciprocity in optomechanical structures. Phys. Rev. Lett., 102, 213903(2009).

    [8] H. Qiu, J. Dong, L. Liu, X. Zhang. Energy-efficient on-chip optical diode based on the optomechanical effect. Opt. Express, 25, 8975-8985(2017).

    [9] M. Hafezi, P. Rabl. Optomechanically induced non-reciprocity in microring resonators. Opt. Express, 20, 7672-7684(2012).

    [10] M. Schmidt, S. Kessler, V. Peano, O. Painter, F. Marquardt. Optomechanical creation of magnetic fields for photons on a lattice. Optica, 2, 635-641(2015).

    [11] A. Metelmann, A. A. Clerk. Nonreciprocal photon transmission and amplification via reservoir engineering. Phys. Rev. X, 5, 021025(2015).

    [12] X.-W. Xu, Y. Li. Optical nonreciprocity and optomechanical circulator in three-mode optomechanical systems. Phys. Rev. A, 91, 053854(2015).

    [13] K. Fang, M. Matheny, X. Luan, O. Painter. Optical transduction and routing of microwave phonons in cavity-optomechanical circuits. Nat. Photonics, 10, 489-496(2016).

    [14] X.-W. Xu, Y. Li, A.-X. Chen, Y.-X. Liu. Nonreciprocal conversion between microwave and optical photons in electro-optomechanical systems. Phys. Rev. A, 93, 023827(2016).

    [15] A. Metelmann, A. A. Clerk. Nonreciprocal quantum interactions and devices via autonomous feedforward. Phys. Rev. A, 95, 013837(2017).

    [16] L. Tian, Z. Li. Nonreciprocal quantum-state conversion between microwave and optical photons. Phys. Rev. A, 96, 013808(2017).

    [17] M.-A. Miri, F. Ruesink, E. Verhagen, A. Alù. Optical nonreciprocity based on optomechanical coupling. Phys. Rev. Appl., 7, 064014(2017).

    [18] G. Li, X. Xiao, Y. Li, X. Wang. Tunable optical nonreciprocity and a phonon-photon router in an optomechanical system with coupled mechanical and optical modes. Phys. Rev. A, 97, 023801(2018).

    [19] J. Kim, M. C. Kuzyk, K. Han, H. Wang, G. Bahl. Non-reciprocal Brillouin scattering induced transparency. Nat. Phys., 11, 275-280(2015).

    [20] C.-H. Dong, Z. Shen, C.-L. Zou, Y.-L. Zhang, W. Fu, G.-C. Guo. Brillouin-scattering-induced transparency and non-reciprocal light storage. Nat. Commun., 6, 6193(2015).

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

    [22] F. Ruesink, M.-A. Miri, A. Alù, E. Verhagen. Nonreciprocity and magnetic-free isolation based on optomechanical interactions. Nat. Commun., 7, 13662(2016).

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

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

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

    [26] N. R. Bernier, L. D. Tóth, A. Koottandavida, M. A. Ioannou, D. Malz, A. Nunnenkamp, A. K. Feofanov, T. J. Kippenberg. Nonreciprocal reconfigurable microwave optomechanical circuit. Nat. Commun., 8, 604(2017).

    [27] G. A. Peterson, F. Lecocq, K. Cicak, R. W. Simmonds, J. Aumentado, J. D. Teufel. Demonstration of efficient nonreciprocity in a microwave optomechanical circuit. Phys. Rev. X, 7, 031001(2017).

    [28] S. Barzanjeh, M. Wulf, M. Peruzzo, M. Kalaee, P. B. Dieterle, O. Painter, J. M. Fink. Mechanical on-chip microwave circulator. Nat. Commun., 8, 953(2017).

    [29] D. Jalas, A. Petrov, M. Eich, W. Freude, S. Fan, Z. Yu, R. Baets, M. Popovic, A. Melloni, J. Joannopoulos, M. Vanwolleghem, C. Doerr, H. Renner. What is–and what is not–an optical isolator. Nat. Photonics, 7, 579-582(2013).

    [30] H. Xie, C.-G. Liao, X. Shang, M.-Y. Ye, X.-M. Lin. Phonon blockade in a quadratically coupled optomechanical system. Phys. Rev. A, 96, 013861(2017).

    [31] X.-W. Xu, H.-Q. Shi, A.-X. Chen, Y.-X. Liu. Cross-correlation between photons and phonons in quadratically coupled optomechanical systems. Phys. Rev. A, 98, 013821(2018).

    [32] Y. Shi, Z. Yu, S. Fan. Limitations of nonlinear optical isolators due to dynamic reciprocity. Nat. Photonics, 9, 388-392(2015).

    [33] R. Huang, A. Miranowicz, J.-Q. Liao, F. Nori, H. Jing. Nonreciprocal photon blockade. Phys. Rev. Lett., 121, 153601(2018).

    [34] G. B. Malykin. The Sagnac effect: correct and incorrect explanations. Phys. Usp., 43, 1229-1252(2000).

    [35] H. Lü, Y. Jiang, Y.-Z. Wang, H. Jing. Optomechanically induced transparency in a spinning resonator. Photon. Res., 5, 367-371(2017).

    [36] H. Jing, H. Lü, S. K. Özdemir, T. Carmon, F. Nori. Nanoparticle sensing with a spinning resonator. Optica, 5, 1424-1430(2018).

    [37] S. Maayani, R. Dahan, Y. Kligerman, E. Moses, A. Hassan, H. Jing, F. Nori, D. Christodoulides, T. Carmon. Flying couplers above spinning resonators generate irreversible refraction. Nature, 558, 569-572(2018).

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

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

    [40] C. Doolin, B. D. Hauer, P. H. Kim, A. J. R. MacDonald, H. Ramp, J. P. Davis. Nonlinear optomechanics in the stationary regime. Phys. Rev. A, 89, 053838(2014).

    [41] G. Brawley, M. Vanner, P. Larsen, S. Schmid, A. Boisen, W. Bowen. Non-linear optomechanical measurement of mechanical motion. Nat. Commun., 7, 10988(2016).

    [42] J. Thompson, B. Zwickl, A. Jayich, F. Marquardt, S. Girvin, J. Harris. Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane. Nature, 452, 72-75(2008).

    [43] H. Jing, D. S. Goldbaum, L. Buchmann, P. Meystre. Quantum optomechanics of a Bose-Einstein antiferromagnet. Phys. Rev. Lett., 106, 223601(2011).

    [44] G. Heinrich, J. G. E. Harris, F. Marquardt. Photon shuttle: Landau-Zener-Stückelberg dynamics in an optomechanical system. Phys. Rev. A, 81, 011801(2010).

    [45] H. Wu, G. Heinrich, F. Marquardt. The effect of Landau-Zener dynamics on phonon lasing. New J. Phys., 15, 123022(2013).

    [46] J. T. Hill. Nonlinear Optics and Wavelength Translation via Cavity-Optomechanics(2013).

    [47] T. K. Paraïso, M. Kalaee, L. Zang, H. Pfeifer, F. Marquardt, O. Painter. Position-squared coupling in a tunable photonic crystal optomechanical cavity. Phys. Rev. X, 5, 041024(2015).

    [48] H.-K. Li, Y.-C. Liu, X. Yi, C.-L. Zou, X.-X. Ren, Y.-F. Xiao. Proposal for a near-field optomechanical system with enhanced linear and quadratic coupling. Phys. Rev. A, 85, 053832(2012).

    [49] H. Wang, Q. Qiao, C. Peng, J. Xia, G. Zhou, Y.-J. Zhao, X.-W. Xu. Two-dimensional optomechanics formed by the graphene sheet and photonic crystal cavity(2018).

    [50] C. W. Gardiner, M. J. Collett. Input and output in damped quantum systems: quantum stochastic differential equations and the master equation. Phys. Rev. A, 31, 3761-3774(1985).

    [51] H. J. Carmichael. An Open Systems Approach to Quantum Optics(1993).

    [52] G. S. Agarwal, S. Huang. Electromagnetically induced transparency in mechanical effects of light. Phys. Rev. A, 81, 041803(2010).

    [53] S. Weis, R. Rivière, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, T. J. Kippenberg. Optomechanically induced transparency. Science, 330, 1520-1523(2010).

    [54] A. H. Safavi-Naeini, T. P. M. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, O. Painter. Electromagnetically induced transparency and slow light with optomechanics. Nature, 472, 69-73(2011).

    [55] S. E. Harris. Electromagnetically induced transparency. Phys. Today, 50, 36-42(1997).

    [56] M. Fleischhauer, A. Imamoglu, J. P. Marangos. Electromagnetically induced transparency: optics in coherent media. Rev. Mod. Phys., 77, 633-673(2005).

    [57] A. Xuereb, C. Genes, A. Dantan. Strong coupling and long-range collective interactions in optomechanical arrays. Phys. Rev. Lett., 109, 223601(2012).

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

    [59] J.-M. Pirkkalainen, S. Cho, F. Massel, J. Tuorila, T. Heikkila, P. Hakonen, M. Sillanpaa. Cavity optomechanics mediated by a quantum two-level system. Nat. Commun., 6, 6981(2015).

    [60] M. J. Burek, J. D. Cohen, S. M. Meenehan, N. El-Sawah, C. Chia, T. Ruelle, S. Meesala, J. Rochman, H. A. Atikian, M. Markham, D. J. Twitchen, M. D. Lukin, O. Painter, M. Lončar. Diamond optomechanical crystals. Optica, 3, 1404-1411(2016).

    [61] H. Zhang, X. Zhao, Y. Wang, Q. Huang, J. Xia. Femtogram scale high frequency nano-optomechanical resonators in water. Opt. Express, 25, 821-830(2017).

    [62] J. D. Cohen, S. M. Meenehan, G. S. MacCabe, S. Gröblacher, A. H. Safavi-Naeini, F. Marsili, M. D. Shaw, O. Painter. Phonon counting and intensity interferometry of a nanomechanical resonator. Nature, 520, 522-525(2015).

    [63] J. Zhang, B. Peng, S. K. Özdemir, Y.-X. Liu, H. Jing, X.-Y. Lü, Y.-L. Liu, L. Yang, F. Nori. Giant nonlinearity via breaking parity-time symmetry: a route to low-threshold phonon diodes. Phys. Rev. B, 92, 115407(2015).

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

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

    [66] H. Wang, Z. Wang, J. Zhang, S. K. Özdemir, L. Yang, Y.-X. Liu. Phonon amplification in two coupled cavities containing one mechanical resonator. Phys. Rev. A, 90, 053814(2014).

    [67] H. Lü, S. K. Özdemir, L.-M. Kuang, F. Nori, H. Jing. Exceptional points in random-defect phonon lasers. Phys. Rev. Appl., 8, 044020(2017).

    [68] Y. Jiang, S. Maayani, T. Carmon, F. Nori, H. Jing. Nonreciprocal phonon laser. Phys. Rev. Appl., 10, 064037(2018).

    [69] Y.-L. Zhang, C.-L. Zou, C.-S. Yang, H. Jing, C.-H. Dong, G.-C. Guo, X.-B. Zou. Phase-controlled phonon laser. New J. Phys., 20, 093005(2018).

    [70] J. Zhang, B. Peng, S. Ozdemir, K. Pichler, D. Krimer, G. Zhao, F. Nori, Y.-X. Liu, S. Rotter, L. Yang. A phonon laser operating at an exceptional point. Nat. Photonics, 12, 479-484(2018).

    [71] S. Hua, J. Wen, X. Jiang, Q. Hua, L. Jiang, M. Xiao. Demonstration of a chip-based optical isolator with parametric amplification. Nat. Commun., 7, 13657(2016).

    [72] K. M. Sliwa, M. Hatridge, A. Narla, S. Shankar, L. Frunzio, R. J. Schoelkopf, M. H. Devoret. Reconfigurable Josephson circulator/directional amplifier. Phys. Rev. X, 5, 041020(2015).

    Xunwei Xu, Yanjun Zhao, Hui Wang, Hui Jing, Aixi Chen. Quantum nonreciprocality in quadratic optomechanics[J]. Photonics Research, 2020, 8(2): 143
    Download Citation