Open Access
22 August 2019 Synchronization and temporal nonreciprocity of optical microresonators via spontaneous symmetry breaking
Da Xu, Zi-Zhao Han, Yu-Kun Lu, Qihuang Gong, Cheng-Wei Qiu, Gang Chen, Yun-Feng Xiao
Author Affiliations +
Abstract

Synchronization is of importance in both fundamental and applied physics, but its demonstration at the micro/nanoscale is mainly limited to low-frequency oscillations such as mechanical resonators. We report the synchronization of two coupled optical microresonators, in which the high-frequency resonances in the optical domain are aligned with reduced noise. It is found that two types of synchronization regimes emerge with either the first- or second-order transition, both presenting a process of spontaneous symmetry breaking. In the second-order regime, the synchronization happens with an invariant topological character number and a larger detuning than that of the first-order case. Furthermore, an unconventional hysteresis behavior is revealed for a time-dependent coupling strength, breaking the static limitation and the temporal reciprocity. The synchronization of optical microresonators offers great potential in reconfigurable simulations of many-body physics and scalable photonic devices on a chip.

1.

Introduction

Synchronization phenomena are ubiquitously observed in nature, such as collective neuron bursts, stabilized heartbeats, and disciplined synchronous fireflies.13 Starting from the Huygens pendulum locked in antiphase,4,5 the synchronization of nonlinear oscillators has earned in-depth investigation.6 In daily life and modern industry, synchronization has been the basis for clock calibration, signal processing, and microwave communication,7 and provides schemes for clustered computing and memory storage.810 Over the past few years, the synchronization of mechanical resonators has been implemented, where the mechanical resonators are coupled strongly through direct conjunction elements,11,12 optical radiation fields,1317 or optical traveling waves,1821 facilitating mechanical-based high performance networks. Strong mutual coupling together with the nonlinearity of individually sustainable systems plays a crucial role in the realization of synchronization.2227

Likewise, synchronized optical fields also promise great potential in fundamental and applied physics, such as many-body optical physics and scalable on-chip photonic devices,2833 while the occurrence is challenged by their relatively low mutual coupling compared to the high carrier frequencies of light. Recently, microcomb solitons have been experimentally synchronized,34,35 significantly expanding their photonic applications, yet the repetition rates in the range of microwaves rather than the optical frequency of the comb lines are equalized. In this article, we discuss the mode synchronization of two optical microresonators without an external reference frequency, where the distant modes are self-sustained and mutually aligned through a weak coupling. The synchronization results from the spontaneous symmetry breaking and takes the form of a first- or second-order transition. Furthermore, an unconventional hysteresis behavior is presented as the coupling strength varies, permitting nonreciprocal synchronization in a more extensive parametric space.

2.

Results

2.1.

Two Self-Sustained Microresonators and Interaction Model

As shown in Fig. 1(a), the system is composed of two optical microcavities with different resonant frequencies, ω10 and ω20, coupled at the strength g. The j’th (j=1,2) cavity is self-sustained by the internal gain described by the factor Gp,j and the intrinsic dissipation at the rate κj. In the presence of nonlinear gain, a self-Kerr type modulation δj(ajaj)2/2 is present, with aj being the annihilation operator and the factor δj describing the self-Kerr effect.36 With the gain saturation37 or the multiphoton absorption,3840 the effective dissipation of the j’th mode is modeled as Kj=κj/2+Rjajaj, where Rj is the nonlinearity factor.

Fig. 1

Schematic diagram of the system. (a) Two detuned and self-sustained optical microcavities with different resonant frequencies, ω10 and ω20, which are directly coupled at strength g. (b)–(d) Frequency spectra of the coupled cavities, showing three different long-term states: unsynchronized, limit cycle (LC), and synchronized (Sync.). Light blue represents the noise backgrounds from which the first- and second-order synchronizations are distinguished.

AP_1_4_046002_f001.png

The dissipative evolution of the system is described by the Lindblad density-matrix equation (=1 hereafter),

Eq. (1)

ρ˙=i[H,ρ]+j=1,2(GjD[aj]ρ+Rj2D[aj2]ρ).
Here, D[o]ρ=2oρoooρρoo and Gj=Gp,jκj/2 denote the net gain factor. Without an external frequency reference, the time-independent Hamiltonian (H)is  equal  toj=1,2[ωj0ajaj+δj(ajaj)2/2]+g(a2a1+a1a2), under the rotating-wave approximation. For simplicity, in the following we set R1=R2=R, G1=G2=G, and δ1=δ2=δ; the dimensionless parameters are defined as δ˜=δ/R, Δ˜=(ω10ω20)/G, and g˜=g/G. The time scale τ=Gt. These formalisms can be checked from wave functions in systems such as coupled laser systems.41,42 Though the coupling between the two modes is linear and energy-conservative, it plays the role of messenger passing over the weak and detuned drive. The self-sustained system always favors the resonance mutual driving, and the synchronization of the two modes is established by the spontaneous frequency alignment of the individual cavities, through the self-Kerr effect and the amplitude stabilization under the saturation effect. In this way, the modes are synchronized in individual cavities.

2.2.

Synchrony Solution in Static Case

We focus on the phase difference and the transient frequencies of two modes in the coherent-state representation.6 In this representation, the complex amplitude αj=aj is parameterized as rjG/Reiϕj, where rj and ϕj are the amplitude and the phase, respectively. Let ϕ=ϕ1ϕ2 be the phase difference, which is the preserved degree of freedom, and let ωj=ϕ˙j be the transient frequency for the j’th mode.

Following the standard Wigner function formalism,43 the mode equation is described by (Λ˙,Λ˙¯)=f(Λ,Λ¯), where Λ=(α1,α2) and f(Λ,Λ¯) denote the quasiprobability drift flow of the two modes (see Supplementary Material for details). The synchrony solution is achieved when f(Λ,Λ¯)=0, and a fixed point Λs(Δ˜,δ˜,g˜) emerges in the parametric space (see Supplementary Material for details). In Fig. 2, we plot the phase differences and the transient frequencies for different g˜. Three categories of long-term behaviors are discovered. When the coupling strength is low, g˜=0.3, for example, the phase difference ϕ accumulates to infinity quickly [see Fig. 2(a1)], and the transient frequencies ω1 and ω2 are effectively separated [Fig. 2(a2)], showing two separated modes in the frequency spectrum [Fig. 1(b)]. When the coupling strength is turned higher, g˜=0.398, for example, the phase difference ϕ vibrates around the stationary point but does not accumulate [Fig. 2(b1)], and the frequencies ω1 and ω2 breathe slowly around the stationary frequency [Fig. 2(b2)], generating a limit cycle state. A stationary mode is localized around the original two cavity modes, and a pair of weak limit cycle modes can be found symmetrically detuned from the stationary modes [Fig. 1(c)]. Finally, with a high enough coupling strength, such as g˜=0.4, the phase difference ϕ stabilizes [Fig. 2(c1)], and the frequencies ω1 and ω2 also converge to a single value [Fig. 2(c2)], reaching the synchronized state with a single mode in the frequency spectrum [Fig. 1(d)].

Fig. 2

Long-term evolutions of the two cavity modes under different coupling strengths. Three different categories are shown: (a) the unsynchronized (g˜=0.3), (b) limit cycle (g˜=0.398), and (c) synchronized states (g˜=0.4). (a1)–(c1) Phase difference; (a2)–(c2) transient frequencies; (a3)–(c3) trajectory encircling types (black cross as the axis); and (a4)–(c4) dynamical potential near the synchrony point. In all figures, the given detuning Δ˜=0.3 and Kerr factor δ˜=0.1.

AP_1_4_046002_f002.png

It is noted that the temporal translational symmetry (TTS) is preserved in the synchronized state because both the amplitudes and phase difference remain invariant while the symmetry is broken in the unsynchronized and limit cycle states. The discrete topological character number as the average encircling number is further defined as

Eq. (2)

χ=T02πlimT|1T0Tdt(ω1ω2)|,
with T0 being the period of long-term evolution (see Supplementary Material for details).44 As shown in Fig. 2(a3), the unsynchronized trajectory encircles the axis r1=r2=0 and has the character number χ=1. The later two categories of trajectories, the off-axial circles [Fig. 2(b3)] and the fixed points [Fig. 2(c3)], have the character number χ=0 (see the transformed space in Supplementary Material for details). With the different symmetries and character numbers, the three long-term states are classified accordingly (see Supplementary Material for details).

2.3.

Analysis of Different Transition Types

When we further study the maximum of the frequency differences, max|ω1ω2|, two types of synchronization transitions are found. As shown in Fig. 3(a), when the coupling strength g˜ is low, the maximal frequency difference varies slowly. At a critical strength g˜c, it suddenly falls to zero, which shows the characteristics of the first-order transition. As shown in Fig. 3(b), the maximal frequency difference continuously decreases to zero but has a discontinuity in its derivative at g˜c, showing the feature of the second-order transition. In addition, the noise spectrum is also calculated in long-term motions. For the synchronized spectrum in Fig. 1(d), the background noise has coinciding peaks with synchronized frequencies in the first-order transition, whereas the noise has shifted-away peaks in the second-order transition (see Supplementary Material for details).

Fig. 3

Parameter dependence of the synchronization. (a), (b) Maximum of the frequency differences, max|ω1ω2|, versus the coupling strength g˜, with (Δ˜=0.2, δ˜=0.1) in (a) and (Δ˜=0.3, δ˜=0.1) in (b); inset shows the derivative. (c) Phase diagram in the (Δ˜,g˜) plane with the Kerr factor δ˜=0.1. The inaccessible (gray), limit cycle (dark blue), and synchronized (light blue) regimes are marked. The red cross stands for the triple phase point (Δ˜T,g˜T). (d) The triple phase point (Δ˜T,g˜T) depending on the Kerr factor δ˜.

AP_1_4_046002_f003.png

In order to study the critical coupling strength g˜c and the transition behaviors in its vicinity, a real-valued dynamical potential V(Λ,Λ¯) is defined (see Supplementary Material for details). Only if the dynamical potential has a local minimum, the fixed point Λs(Δ˜,δ˜,g˜) emerges and remains stable, and thus indicates the existence of a synchronized state (see Supplementary Material for details). In the vicinity of the fixed point, the dynamical potential can be expanded as V(Λ,Λ¯)=V(Λs,Λ¯s)12[(ΔΛ¯,ΔΛ)·J·(ΔΛ¯,ΔΛ)+H.c.], where J(Λ,Λ¯)=f(Λ,Λ¯)/(Λ,Λ¯) is the Jacobian matrix, ΔΛ is the arbitrarily small displacement from the fixed point, and H.c. is the Hermitian conjugate. The displacement ΔΛ signifies the breaking of the TTS. At ΔΛ=0, the TTS is preserved. The stability near the fixed point is thus governed by the eigenvalues of the Jacobian. When the largest real part of the J eigenvalues [known as the largest Lyapunov exponent L(Λs)] is positive, the fixed point is unstable and vice versa (see Supplementary Material for details).45 The critical coupling strength g˜c is then taken at L(Λs)=0.

In the three-dimensional space (r1,r2,ϕ), the Jacobian J has purely real 3×3 components, and thus the complex eigenvalues must come in pairs. If the largest Lyapunov exponent L equals one of the eigenvalues, the dynamical potential is simplified as

Eq. (3)

V(x)=b0(g˜g˜c)x2,
where x is the perturbation of Λs in the direction of corresponding eigenvector, and the real coefficient b0=dL/dg˜>0 (see Supplementary Material for details). It is noted that the dynamical potential in Eq. (3) becomes a well or a barrier depending on g˜>g˜c or g˜<g˜c, which leads to the synchronized or unsynchronized state shown in Figs. 2(a4) and 2(c4). Thus, the first-order transition happens at g˜c, explaining the sudden convergence of frequency difference in Fig. 3(a). For g˜<g˜c and g˜>g˜c, the TTS is broken (ΔΛ) and preserved (ΔΛ=0), respectively. If the largest Lyapunov exponent L equals the real parts of a pair of conjugating eigenvalues, the averaged dynamical potential is

Eq. (4)

V=b1(g˜g˜c)ρ2+b2ρ4,
where ρ is the radial displacement from Λs and the real coefficients {b1,b2}>0 (see Supplementary Material for details).46 When g˜<g˜c, a double-well type potential is obtained, corresponding to the limit cycle state shown in Fig. 2(b4). After g˜ surpasses g˜c, the averaged dynamical potential V has a single local minimum at Λs, and the synchronization is reached, accounting for the second-order transition depicted in Fig. 3(b). The TTS is spontaneously broken as the radial displacement ρ continuously departs from the synchrony point Λs.

In the light of the static analysis above, the phase diagram in the δ˜-cross section is plotted in Fig. 3(c), where three regions of different long-term behaviors are marked. The synchronized and limit cycle regimes are specified according to the existence of a single and double local minima of the dynamical potentials, respectively. The inaccessible (unsynchronized) regime corresponds to the saddle nodes in the dynamical potentials, and thus neither the synchronized state nor the limit cycle state can survive in this regime. The transition from the unsynchronized state to the synchronized state is of first-order and has a variant topological character number. The transition from the limit cycle state to the synchronized state is of second-order and has an invariant topological character number (see Supplementary Material for details). It is also found that a triple phase point emerges at g˜c=g˜T and Δ˜=Δ˜T, where Δ˜T is the minimal detuning required for the second-order transition. This point corresponds to the solution where two eigenvalues of the Jacobi matrix J are zeros. For the detuning Δ˜<Δ˜T (Δ˜>Δ˜T), the first-order (second-order) transition happens around g˜c (black solid line). The triple phase point relies crucially on the strength of the Kerr effect. In Fig. 3(d), we plot Δ˜T and g˜T with respect to the Kerr factor δ˜, showing monotone increasing and decreasing dependence, respectively. When the factor δ˜ increases, the self-tuning ability of the Kerr effect is strengthened, and the second-order synchronization under a larger detuning and a weaker coupling becomes possible.

2.4.

Hysteresis Behavior

The two types of synchronization transitions present distinct hysteresis behaviors near the critical coupling strength g˜c. In Fig. 4, the frequency differences |ω1ω2| and their maxima, max|ω1ω2|, are plotted, to indicate when the real-time coupling strength g˜(τ) slowly increases (forward) and then decreases (backward). In the first-order transition regime, whatever direction g˜(τ) moves, the synchronization emerges or disappears at the same g˜c as derived in the static analysis [see Fig. 4(a)]. The maximal frequency differences in the forward and backward evolutions are identical to those depicted in Fig. 4(c), coinciding with Fig. 3(a). In the second-order transition regime, although the synchronization also emerges at g˜c in the forward trip, it does not disappear at the same critical point in the backward trip [see Fig. 4(b)]. Actually, the synchronization survives far below g˜c, even into the statically inaccessible region depicted in Fig. 3(c). Further calculation of the maximal frequency differences reveals a hysteresis loop in this case [see Fig. 4(d)]. The forward half of the loop remains the same as the curve in Fig. 3(b), whereas the backward half is beneath it. The critical coupling strength under the static model does not apply under a dynamical model with the second-order transition. As explained in the Supplementary Material,47 the g˜(τ) passes g˜c with the emergence of new non-zero eigenvalues proportional to g˜˙J1. This mechanism is thus attributed to the singularity of J1 and the altering direction of real-time g˜(τ). The hysteresis property breaks the minimal coupling required for the synchronization, and the consequent temporal nonreciprocity enables the reading out of coupling history as has been done in the ferromagnetic materials.48

Fig. 4

Hysteresis behavior in frequency difference. (a), (b) Frequency differences |ω1ω2| versus the evolution time τ in the first- and second-order transition regimes. Insets: the real-time evolution of the coupling strength g˜(τ). (c), (d) Maxima of the frequency differences, max|ω1ω2| versus g˜(τ). For each plot, the Kerr factor δ˜=0.1; the detuning Δ˜=0.2 in (a) and (c), and Δ˜=0.3 in (b) and (d).

AP_1_4_046002_f004.png

3.

Discussion and Conclusion

In summary, we have presented the mode synchronization of two self-sustained optical microresonators that are largely detuned and linearly coupled together. The synchronization is accompanied by a process of spontaneous symmetry breaking, taking the form of the first- and second-order transitions. First, when the synchronization takes place, the high transient frequencies of both modes collapse, offering a possible solution to the frequency mismatch problem in integrating optical microresonators. The phase noise of the coupled system is dramatically reduced, revealing spontaneous symmetry preservation and paving the way for error-tolerant device fabrication.19 Second, the topological character transitions cast light on many-body physics. The experimental realization can be approached by coupling two toroid cavities etched with the same mask. Raman gain is applied separately to each cavity at tunable pump frequencies. With additional thermal control of the refractive index, perfect phase matching and adjustable mode frequency differences are achievable. Note that in our model the evolution in the transformed space corresponds to the nontrivial degeneration of a ring into a point. During the synchronization of three resonators, however, the transition also includes nontrivial degeneration of the torus into a ring. The multiple-torus topological structure in massively coupled resonators offers new insights for many-body physics.49 Finally, in the second-order transition regime, an unconventional hysteresis behavior was predicted, breaking the static critical coupling strength limit. The coupling history of the resonators can be logged over a short period autonomously, which is desirable in all-optical memory designs.50,51 These results thus show great potential for further research studies in all-optical memory, coupled cavity quantum electrodynamics, and many-body optical physics.

Acknowledgments

We thank Linran Fan, Qi-Tao Cao, and Mian Zhang for fruitful discussions. This work was supported by the National Key R&D Program of China (Grant Nos. 2016YFA0301302 and 2018YFB2200401), NSFC (Grant Nos. 11825402, 61435001, 11654003, and 11674200), and High-Performance Computing Platform of Peking University. The authors declare that they have no competing financial interests.

References

1. 

B. van der Pol and J. van der Mark, “LXXII. The heartbeat considered as a relaxation oscillation, and an electrical model of the heart,” London Edinburgh Dublin Philos. Mag. J. Sci., 6 (38), 763 –775 (2009). https://doi.org/10.1080/14786441108564652 Google Scholar

2. 

R. FitzHugh, “Impulses and physiological states in theoretical models of nerve membrane,” Biophys. J., 1 (6), 445 –466 (1961). https://doi.org/10.1016/S0006-3495(61)86902-6 BIOJAU 0006-3495 Google Scholar

3. 

J. Buck and E. Buck, “Mechanism of rhythmic synchronous flashing of fireflies: fireflies of Southeast Asia may use anticipatory time-measuring in synchronizing their flashing,” Science, 159 (3821), 1319 –1327 (1968). https://doi.org/10.1126/science.159.3821.1319 SCIEAS 0036-8075 Google Scholar

4. 

C. Huygens, Oeuvres Complètes, 7 M. Nijhoff, Leiden (1897). Google Scholar

5. 

H. M. Oliveira and L. V. Melo, “Huygens synchronization of two clocks,” Sci. Rep., 5 11548 (2015). https://doi.org/10.1038/srep11548 SRCEC3 2045-2322 Google Scholar

6. 

J. Kurths, A. Pikovsky and M. Rosenblum, Synchronization: A Universal Concept in Nonlinear Sciences, Cambridge University Press, New York (2001). Google Scholar

7. 

S. Bregni, Synchronization of Digital Telecommunications Networks, 27 Wiley, New York (2002). Google Scholar

8. 

M. Bagheri et al., “Dynamic manipulation of nanomechanical resonators in the high-amplitude regime and non-volatile mechanical memory operation,” Nat. Nanotechnol., 6 (11), 726 –732 (2011). https://doi.org/10.1038/nnano.2011.180 NNAABX 1748-3387 Google Scholar

9. 

I. Mahboob and H. Yamaguchi, “Bit storage and bit flip operations in an electromechanical oscillator,” Nat. Nanotechnol., 3 (5), 275 –279 (2008). https://doi.org/10.1038/nnano.2008.84 NNAABX 1748-3387 Google Scholar

10. 

F. C. Hoppensteadt and E. M. Izhikevich, “Synchronization of MEMS resonators and mechanical neurocomputing,” IEEE Trans. Circuits Syst. I, 48 (2), 133 –138 (2001). https://doi.org/10.1109/81.904877 ITCAEX 1057-7122 Google Scholar

11. 

H. Nijmeijer and A. Rodriguez-Angeles, Synchronization of Mechanical Systems, 46 World Scientific, Singapore (2003). Google Scholar

12. 

S.-B. Shim, M. Imboden and P. Mohanty, “Synchronized oscillation in coupled nanomechanical oscillators,” Science, 316 (5821), 95 –99 (2007). https://doi.org/10.1126/science.1137307 SCIEAS 0036-8075 Google Scholar

13. 

G. Heinrich et al., “Collective dynamics in optomechanical arrays,” Phys. Rev. Lett., 107 043603 (2011). https://doi.org/10.1103/PhysRevLett.107.043603 PRLTAO 0031-9007 Google Scholar

14. 

C. A. Holmes, C. P. Meaney and G. J. Milburn, “Synchronization of many nanomechanical resonators coupled via a common cavity field,” Phys. Rev. E, 85 066203 (2012). https://doi.org/10.1103/PhysRevE.85.066203 Google Scholar

15. 

M. Zhang et al., “Synchronization of micromechanical oscillators using light,” Phys. Rev. Lett., 109 (23), 233906 (2012). https://doi.org/10.1103/PhysRevLett.109.233906 PRLTAO 0031-9007 Google Scholar

16. 

M. Zhang et al., “Synchronization and phase noise reduction in micromechanical oscillator arrays coupled through light,” Phys. Rev. Lett., 115 (16), 163902 (2015). https://doi.org/10.1103/PhysRevLett.115.163902 PRLTAO 0031-9007 Google Scholar

17. 

V. Peano et al., “Topological phases of sound and light,” Phys. Rev. X, 5 (3), 031011 (2015). https://doi.org/10.1103/PhysRevX.5.031011 PRXHAE 2160-3308 Google Scholar

18. 

S. Y. Shah et al., “Master-slave locking of optomechanical oscillators over a long distance,” Phys. Rev. Lett., 114 113602 (2015). https://doi.org/10.1103/PhysRevLett.114.113602 PRLTAO 0031-9007 Google Scholar

19. 

M. Bagheri et al., “Photonic cavity synchronization of nanomechanical oscillators,” Phys. Rev. Lett., 111 (21), 213902 (2013). https://doi.org/10.1103/PhysRevLett.111.213902 PRLTAO 0031-9007 Google Scholar

20. 

T. Li et al., “Long-distance synchronization of unidirectionally cascaded optomechanical systems,” Opt. Express, 24 (11), 12336 –12348 (2016). https://doi.org/10.1364/OE.24.012336 OPEXFF 1094-4087 Google Scholar

21. 

E. Gil-Santos et al., “Light-mediated cascaded locking of multiple nano-optomechanical oscillators,” Phys. Rev. Lett., 118 063605 (2017). https://doi.org/10.1103/PhysRevLett.118.063605 PRLTAO 0031-9007 Google Scholar

22. 

M. Cross et al., “Synchronization by nonlinear frequency pulling,” Phys. Rev. Lett., 93 (22), 224101 (2004). https://doi.org/10.1103/PhysRevLett.93.224101 PRLTAO 0031-9007 Google Scholar

23. 

D. K. Agrawal, J. Woodhouse and A. A. Seshia, “Observation of locked phase dynamics and enhanced frequency stability in synchronized micromechanical oscillators,” Phys. Rev. Lett., 111 084101 (2013). https://doi.org/10.1103/PhysRevLett.111.084101 PRLTAO 0031-9007 Google Scholar

24. 

S. Walter, A. Nunnenkamp and C. Bruder, “Quantum synchronization of a driven self-sustained oscillator,” Phys. Rev. Lett., 112 (9), 094102 (2014). https://doi.org/10.1103/PhysRevLett.112.094102 PRLTAO 0031-9007 Google Scholar

25. 

L. M. Pecora et al., “Cluster synchronization and isolated desynchronization in complex networks with symmetries,” Nat. Commun., 5 4079 (2014). https://doi.org/10.1038/ncomms5079 NCAOBW 2041-1723 Google Scholar

26. 

M. H. Matheny et al., “Phase synchronization of two anharmonic nanomechanical oscillators,” Phys. Rev. Lett., 112 (1), 014101 (2014). https://doi.org/10.1103/PhysRevLett.112.014101 Google Scholar

27. 

N. Lörch et al., “Genuine quantum signatures in synchronization of anharmonic self-oscillators,” Phys. Rev. Lett., 117 (7), 073601 (2016). https://doi.org/10.1103/PhysRevLett.117.073601 PRLTAO 0031-9007 Google Scholar

28. 

E. Kuramochi et al., “Large-scale integration of wavelength-addressable all-optical memories on a photonic crystal chip,” Nat. Photonics, 8 (6), 474 –481 (2014). https://doi.org/10.1038/nphoton.2014.93 NPAHBY 1749-4885 Google Scholar

29. 

M. Zhang et al., “Electronically programmable photonic molecule,” Nat. Photonics, 13 (1), 36 –40 (2019). https://doi.org/10.1038/s41566-018-0317-y NPAHBY 1749-4885 Google Scholar

30. 

Y.-C. Liu et al., “Coherent polariton dynamics in coupled highly dissipative cavities,” Phys. Rev. Lett., 112 (21), 213602 (2014). https://doi.org/10.1103/PhysRevLett.112.213602 PRLTAO 0031-9007 Google Scholar

31. 

M.-J. Hwang and M. B. Plenio, “Quantum phase transition in the finite Jaynes-Cummings lattice systems,” Phys. Rev. Lett., 117 (12), 123602 (2016). https://doi.org/10.1103/PhysRevLett.117.123602 PRLTAO 0031-9007 Google Scholar

32. 

J. Tangpanitanon et al., “Topological pumping of photons in nonlinear resonator arrays,” Phys. Rev. Lett., 117 (21), 213603 (2016). https://doi.org/10.1103/PhysRevLett.117.213603 PRLTAO 0031-9007 Google Scholar

33. 

M. Ludwig and F. Marquardt, “Quantum many-body dynamics in optomechanical arrays,” Phys. Rev. Lett., 111 (7), 073603 (2013). https://doi.org/10.1103/PhysRevLett.111.073603 PRLTAO 0031-9007 Google Scholar

34. 

J. K. Jang et al., “Synchronization of coupled optical microresonators,” Nat. Photonics, 12 (11), 688 –693 (2018). https://doi.org/10.1038/s41566-018-0261-x NPAHBY 1749-4885 Google Scholar

35. 

Q.-F. Yang et al., “Counter-propagating solitons in microresonators,” Nat. Photonics, 11 (9), 560 –564 (2017). https://doi.org/10.1038/nphoton.2017.117 Google Scholar

36. 

C. Jirauschek and F. X. Kärtner, “Gaussian pulse dynamics in gain media with Kerr nonlinearity,” J. Opt. Soc. Am. B, 23 (9), 1776 –1784 (2006). https://doi.org/10.1364/JOSAB.23.001776 JOBPDE 0740-3224 Google Scholar

37. 

D. Liu et al., “Symmetry, stability, and computation of degenerate lasing modes,” Phys. Rev. A, 95 (2), 023835 (2017). https://doi.org/10.1103/PhysRevA.95.023835 Google Scholar

38. 

T. E. Lee and H. Sadeghpour, “Quantum synchronization of quantum van der Pol oscillators with trapped ions,” Phys. Rev. Lett., 111 (23), 234101 (2013). https://doi.org/10.1103/PhysRevLett.111.234101 PRLTAO 0031-9007 Google Scholar

39. 

D. F. Walls and G. J. Milburn, Quantum Optics, Springer Science & Business Media, Berlin (2007). Google Scholar

40. 

L. Gilles and P. Knight, “Two-photon absorption and nonclassical states of light,” Phys. Rev. A, 48 (2), 1582 –1593 (1993). https://doi.org/10.1103/PhysRevA.48.1582 Google Scholar

41. 

S. Wieczorek and W. W. Chow, “Bifurcations and interacting modes in coupled lasers: a strong-coupling theory,” Phys. Rev. A, 69 033811 (2004). https://doi.org/10.1103/PhysRevA.69.033811 Google Scholar

42. 

S. Wieczorek and W. W. Chow, “Chaos in practically isolated microcavity lasers,” Phys. Rev. Lett., 92 (21), 213901 (2004). https://doi.org/10.1103/PhysRevLett.92.213901 PRLTAO 0031-9007 Google Scholar

43. 

H. Carmichael, Statistical Methods in Quantum Optics 1: Master Equations and Fokker-Planck Equations, Springer, Berlin (1999). Google Scholar

44. 

A. Mari et al., “Measures of quantum synchronization in continuous variable systems,” Phys. Rev. Lett., 111 (10), 103605 (2013). https://doi.org/10.1103/PhysRevLett.111.103605 PRLTAO 0031-9007 Google Scholar

45. 

A. M. Lyapunov, “The general problem of the stability of motion,” Int. J. Control, 55 (3), 531 –534 (1992). https://doi.org/10.1080/00207179208934253 Google Scholar

46. 

J. E. Marsden and M. McCracken, The Hopf Bifurcation and Its Applications, 19 Springer Science & Business Media, Heildelberg (2012). Google Scholar

47. 

N. Berglund, “Dynamic bifurcations: hysteresis, scaling laws and feedback control,” Prog. Theor. Phys. Suppl., 139 325 –336 (2000). https://doi.org/10.1143/PTPS.139.325 Google Scholar

48. 

K. H. J. Buschow and F. R. Boer, Physics of Magnetism and Magnetic Materials, 92 Springer, New York (2003). Google Scholar

49. 

M. A. Armstrong, Basic Topology, Springer Science & Business Media, Heildelberg (2013). Google Scholar

50. 

C.-H. Chen et al., “All-optical memory based on injection-locking bistability in photonic crystal lasers,” Opt. Express, 19 (4), 3387 –3395 (2011). https://doi.org/10.1364/OE.19.003387 OPEXFF 1094-4087 Google Scholar

51. 

C. Ríos et al., “Integrated all-photonic non-volatile multi-level memory,” Nat. Photonics, 9 (11), 725 –732 (2015). https://doi.org/10.1038/nphoton.2015.182 NPAHBY 1749-4885 Google Scholar

Biography

Da Xu received his BS degree (fundamental science of mathematics and physics), in 2016 from Tsinghua University. Currently, he serves as a PhD student in the State Key Laboratory for Mesoscopic Physics and School of Physics, Peking University. His research interests include quantum optics, optomechanics, and microcavity photonics.

Gang Chen received his BS and PhD degrees in physics from Shaoxing University and Shanxi University, in 2001 and 2009, respectively. After getting his PhD, he joined the faculty of Shanxi University and was promoted to full professor in 2012. His research interests lie in the fields of ultracold atoms and quantum optics. He has been awarded funding by the National Science Foundation for Excellent Young Scholars of China.

Yun-Feng Xiao received his BS and PhD degrees in physics from University of Science and Technology of China in 2002 and 2007, respectively. After a postdoctoral research at Washington University in St. Louis, he joined the faculty of Peking University in January of 2009. He was promoted to tenured associate in 2014, and full professor in 2018. He was elected as an OSA fellow in 2018. His research interests lie in the fields of whispering-gallery microcavity optics and photonics. He was awarded the Wang Daheng Optics Prize in 2018, the Rao Yutai Prize in fundamental optics in 2013, and the Rao Yutai Prize in physics in 2019.

Biographies of the other authors not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Da Xu, Zi-Zhao Han, Yu-Kun Lu, Qihuang Gong, Cheng-Wei Qiu, Gang Chen, and Yun-Feng Xiao "Synchronization and temporal nonreciprocity of optical microresonators via spontaneous symmetry breaking," Advanced Photonics 1(4), 046002 (22 August 2019). https://doi.org/10.1117/1.AP.1.4.046002
Received: 15 April 2019; Accepted: 29 July 2019; Published: 22 August 2019
Lens.org Logo
CITATIONS
Cited by 13 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Microresonators

Physics

Resonators

Applied physics

Microwave radiation

Optical microcavities

Photonic devices

Back to Top