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

Bloch oscillation and unidirectional translation of frequency in a dynamically modulated ring resonator

Open Access Open Access

Abstract

The ability to control the spectrum of light is of fundamental significance in many applications of light. We consider a dynamically modulated ring resonator that supports a set of resonant modes equally spaced in their resonant frequencies, and is modulated at a frequency that is slightly detuned from the modal frequency spacing. We find that such a system can be mapped into a tight-binding model of photons under a constant effective force, and, as a result, the system exhibits Bloch oscillation along the frequency axis. The sign of the detuning is mapped to the sign of the effective force and hence controls the direction of the Bloch oscillation. We also show that a periodic switching of the detuning can lead to unidirectional transport of light along the frequency axis. Our work points to a new capability for manipulating the frequencies of light using dynamic microcavity structures.

© 2016 Optical Society of America

1. INTRODUCTION

Microring resonators have been extensively used for explorations of fundamental physics and in a wide range of applications including spectroscopy, sensing, and information processing [15]. In particular, the resonant modes of a ring cavity form a frequency comb, and hence the ring can be used as the basis for compact and robust frequency comb sources [68]. In addition, active modulation of microring cavities can couple these resonant modes together, leading to new possibilities for manipulating the spectrum of light [911].

In the absence of group velocity dispersion in the ring waveguide, a ring cavity supports an equally spaced frequency comb [6]. In the on-resonance modulation case, in which the modulation frequency Ω is chosen to be equal to the frequency spacing ΩR between the resonant modes in the comb, the system exhibits photonic transition [1214] and can be mapped into a tight-binding model with each mode corresponding to one lattice site in the model [15,16]. Building upon this notion, in this paper we consider the off-resonance modulation case, in which the modulation frequency is slightly different from the frequency spacing. We show that, in such a case, the system can be mapped into a tight-binding model subject to a constant effective force for photons. Thus, the photon wave amplitudes in the system undergo Bloch oscillation along the frequency axis. Moreover, we find that a unidirectional transport of the light along the frequency axis can be achieved by periodically changing the modulation frequency Ω.

The Bloch oscillation of light has been previously explored in waveguide or waveguide array systems, with the oscillation occurring either in the spatial [1721] or the spectral domain [22,23]. Our work here points to a novel manifestation of the effect of Bloch oscillation [24,25] and establishes a connection between the physics of microresonators and the physics of coherent dynamics of quantum particles under an external field. Moreover, the ability to achieve unidirectional transport of light along the frequency axis naturally leads to the capability of photon frequency translation, which is of interest for both classical and quantum information processing [2628].

2. THEORETICAL ANALYSIS

To illustrate the essential physics, we first consider a simple physical model of a ring resonator undergoing dynamic refractive index modulation (Fig. 1). Suppose the waveguide forming the ring has no group velocity dispersion; the frequency of the mth resonant mode is then located at

ωm=ω0+mΩR,
where ΩR is the free spectral range. Suppose we place an electro-optic modulator into the ring, with a modulation frequency ΩΩR. A simple Hamiltonian for this system is then
H=mωmamam+m2gcos(Ωt)(amam+1+am+1am),
where am and am are the creation and annihilation operators, respectively, for the mth resonant mode in the resonator. g is the strength of modulation. We assume that the modes are all nondegenerate, and there is no backscattering in the waveguide forming the ring. Defining cmameiωmt and taking the rotating wave approximation, which is valid since ΩΩR, we transform H in Eq. (2) to
HRWA=gm(eiΔtcmcm+1+eiΔtcm+1cm),
where ΔΩΩR is the detuning. Equation (3) is equivalent to a time-independent Hamiltonian describing a photon in a one-dimensional lattice under a constant effective force:
H˜RWA=gm(cmcm+1+cm+1cm)+mmΔcmcm,
as can be seen with the gauge transformation [29]
|Ψ=mvmcm|0|Ψ˜=mv˜mcm|0=mvmeimΔtcm|0,
with vm(v˜m) being the amplitude of the field at the mth mode. |Ψ(Ψ˜) satisfies the Schrödinger equation, i(/t)|Ψ(Ψ˜)=HRWA(H˜RWA)|Ψ(Ψ˜). In Eq. (4), we note that the frequency of the mode at the mth site is proportional to m, signifying the presence of a constant force, as schematically represented in Fig. 1(c). The strength of the effective force is F=|Δ/ΩR|, which leads to a Bloch oscillation with a period TB=2π/Δ [24,30].

 figure: Fig. 1.

Fig. 1. (a) Ring resonator modulated by an electro-optic modulator (EOM). (b) In the absence of group velocity dispersion, the ring in (a) supports a set of resonant modes (blue dashed lines) with their resonant frequencies equally spaced by ΩR. The modulation frequency Ω is chosen to be slightly different from ΩR. For the modes that are located at ω0, the modulation generates sidebands that are detuned from the other resonant modes. (c) The system as described in (b) can be mapped onto a tight-binding model of a photon under an effective force [Eq. (4)].

Download Full Size | PDF

We now proceed with a more realistic model of a ring under dynamic modulation [16,31]. We expand the electric field inside the resonator as [32]

E(t,r,z)=mEm(t,z)Em(r)eiωmt,
where z denotes the propagation direction along the waveguide that forms the ring resonator, r gives the directions perpendicular to z, and Em(r) is the modal profile. Em(t,z) is the modal amplitude for the mth resonant mode, which satisfies the following equation:
(z+iβ(ωm))Em(t,z)ng(ωm)ctEm(t,z)=0,
as can be derived from Maxwell’s equations under the slowly varying envelope approximation. Here β is the wavevector in the z direction and ng is the group index. The geometry of the ring imposes a periodic boundary condition Em(t,z+L)=Em(t,z), where L is the circumference of the ring.

The electric field inside the resonator undergoes a dynamic phase modulation when it passes through an electro-optic phase modulator located at z0, as described by

E(t+,r,z0)=E(t,r,z0)eiαcos(Ωt),
where α is the modulation amplitude and t±=t+0±. The modulator imposes an additional time-dependent phase as the field passes through the modulator. We then use the rotating wave approximation to describe this process in terms of Em(t,z) and obtain [33]
Em(t+,z0)=J0(α)Em(t,z0)+qiqJq(α)[Em+q(t,z0)eiqΔt+Emq(t,z0)eiqΔt],
where Δ=ΩΩR and Jq is the Bessel function of the qth order.

To relate the physical model here to the simple model of Eq. (2), we keep only the coupling between nearest neighbor sidebands in Eq. (9), (i.e., keep only q=±1 terms), and also assume that J0(α)1 and J1(α)α/2, which are valid in the weak modulation limit α0. Therefore, the variation of Em in Eq. (9) after it passes the modulator is

ΔEm(t,z0)=iα2[Em+1(t,z0)eiΔt+Em1(t,z0)eiΔt].

The total change of Em over the propagation over one loop is Em(t+TR)=Em(t)+ΔEm(t), where TR is the round-trip propagation time around the ring [34]. By defining Em(t+TR)Em(t)TREm(T)T, we obtain

TREmT=iα2(Em+1eiΔT+Em1eiΔT).

We therefore see that, in the appropriate limit, Eqs. (7) and (9) describe a dynamics that is the same as the dynamics of the Hamiltonian of Eq. (2). Therefore, a ring under modulation can indeed exhibit Bloch oscillation physics in the frequency domain as described in Eq. (2).

3. NUMERICAL DEMONSTRATIONS OF BLOCH OSCILLATION AND TRANSLATION OF FREQUENCY

As a numerical confirmation of the analytical argument above, we solve Eqs. (7) and (9) to obtain the solution of the propagation of the electric field inside a ring resonator modulated by an electro-optic modulator (with α=0.2), without making the approximation that led to Eq. (10) [16]. The simulation involves solving for the dynamics for a finite set of modes around ω0 as labeled by m. Let n0=n(ωm) denote the effective refractive index for all resonant modes, so we have ΩR=2π c/n0L. We inject the excitation into the system at the position zs by the equation

Em(t+,zs)=Em(t,zs)+Sm(t).

Here we assume that the excitation source is inside the ring. Throughout the paper, we choose

Sm(t)=fme2ln2(t20)2/100,
where t is in the unit of n0L/c. Here Sm is a pulse with its frequency centered at ωm and its full width at half-maximum (FWHM) ΔωΩR. fm is the peak excitation amplitude for the mth mode. A given Sm essentially excites only the mth mode.

In the first set of simulations, for the excitation, we choose f0=1 and fm0=0 in Eq. (13). We consider two modulation frequencies corresponding to detunings Δ=0.1 c/n0L and Δ=0.5 c/n0L. In the simulation we include 17 resonant modes labeled with m=8,,8. We plot in Fig. 2 the temporal evolution of the total energy for each resonant mode (Im(t)ring|Em(t,z)|2dz). With a small detuning of Δ=0.1 c/n0L, we see the regular Bloch oscillation in the frequency dimension with the period TB=20π n0L/c. When the detuning is large (Δ=0.5 c/n0L), we no longer observe the spectral Bloch oscillation because the modulation cannot couple the nearby resonant modes efficiently.

 figure: Fig. 2.

Fig. 2. Evolution in time of the total energy for each resonant mode (Im(t)ring|Em(t,z)|2dz): (a) Δ=0.1 c/n0L, and (b) Δ=0.5 c/n0L, when only the m=0 mode is excited.

Download Full Size | PDF

In Fig. 2(a), we observe that the oscillation pattern is symmetric around the m=0 axis, in spite of the fact that the choice of Δ0 gives a directional effective force along the frequency axis, as seen in Eq. (4). This effect was analytically studied in [35]. Physically, the excitation of a single site corresponds to a uniform excitation over the entire band in kf-space (where kf is the wavevector reciprocal to the frequency axis). The application of a force shifts the entire band and, hence, cannot create an asymmetric oscillation.

To observe an asymmetric Bloch oscillation pattern, we now consider an excitation profile by setting

fm=exp[(ωmωc)2/(5ΩR)2],
in Eq. (13) with ωc=ω0. We again perform the simulation including 17 resonant modes. We consider the cases with the positive detuning Δ=0.055 c/n0L and the negative detuning Δ=0.055 c/n0L, respectively. The total energy for each mode Im(t) is shown in Fig. 3. Unlike the case of Fig. 2, here the oscillation pattern no longer has a mirror symmetry around the m=0 axis. The Gaussian excitation profile [Eq. (14)] excites only a portion of the band around kf=0. The effective force can then directionally shift the field amplitude profile in the kf space as well as the frequency space. Changing the sign of the detuning Δ flips the oscillation pattern with respect to the m=0 axis.

 figure: Fig. 3.

Fig. 3. Evolution in time of the total energy for each resonant mode (Im(t)ring|Em(t,z)|2dz): (a) Δ=0.055 c/n0L, and (b) Δ=0.055 c/n0L, when several modes are initially excited.

Download Full Size | PDF

In Fig. 3, we see that, under a constant effective force corresponding to a constant detuning Δ, the total intensity oscillates around the m=0 point in the frequency axis. Next, we show that by varying the direction of the effective force periodically in time, as can be achieved by varying the detuning Δ periodically in time, we can achieve a unidirectional transport of light in the frequency space. In the simulation, we periodically switch Δ between ±0.055 c/n0L at every time interval of t=55 n0L/cπ/|Δ| [see Fig. 4(b)]. In the simulation we include 29 resonant modes (m=14,,14), to accommodate the range of frequency shift in the unidirectional transport process. The excitation follows Eqs. (13) and (14) with ωc=ω12. In Fig. 4(a), we plot the total energy for each mode (Im(t)). The Bloch oscillation is interrupted every time a sudden change in the modulation frequencies occurs, resulting in a unidirectional shift in the frequency space as a function of time.

 figure: Fig. 4.

Fig. 4. (a) The evolution in time of the total energy for each resonant mode (Im(t)), as the modulation frequency is periodically switched between Δ=±0.055 c/n0L, as described in (b).

Download Full Size | PDF

In experiments, a ring is typically excited through the coupling to external waveguides. Here we demonstrate unidirectional frequency transport in the presence of external coupling waveguides (Fig. 5). We assume that the coupling occurs at a position at z1,2 on the ring circumference with the external waveguide 1 or 2, as described by

Em(t+,z1,2)=1γ1,22(t)Em(t,z1,2)iγ1,2(t)Em(1,2)(t,z1,2),
Em(1,2)(t+,z1,2)=1γ1,22(t)Em(1,2)(t,z1,2)iγ1,2(t)Em(t,z1,2),
where Em(1,2) is the amplitude of the field component at the frequency ωm in external waveguide 1 or 2. γ1,2(t) is the coupling strength between the ring resonator and waveguide 1 or 2. For optimal observation of unidirectional transport along the frequency axis, it is preferable that the cavity does not leak into the waveguide during the modulation process. We therefore choose a time-dependent waveguide–cavity coupling constant in the form γ1(t)=0.1H(40t) and γ2(t)=0.1H(t200) in the simulation, where H(t) is the Heaviside step function. Such an on–off behavior in coupling can be achieved with electro-optic tuning of waveguide coupler [32]. With such a choice, the input source is injected into the ring from waveguide 1 at t<40 n0L/c, and, after modulation, the field in the ring leaks out into waveguide 2 at t>200 n0L/c and gets detected. We first perform the simulation with the input field into external waveguide 1 obeying Eqs. (13) and (14) with ωc=ω12. We turn on the dynamic modulation at t=0 with the same periodic switch of the modulation frequency following Fig. 4(b), but turn it off at various times tf (tf=50,100,150,200 n0L/c). In Fig. 5(b), we plot the output spectrum from external waveguide 2 after the modulation is switched off. The top panel of Fig. 5(b) shows the profile of the input spectrum. Compared to it, the output spectra shift to higher frequencies. As tf gets larger, the magnitude of the shift becomes larger, demonstrating the unidirectional frequency shift in this system. In contrast, in Fig. 5(c), we choose the same input field except ωc=ω0 in Eq. (14), and we apply a constant modulation frequency with Δ=0.055 c/n0L instead over the same periods from 0 to the various tf, as discussed above. For all values of tf, the output spectra are located somewhat near ω0, and do not exhibit the unidirectional shift as we observe in Fig. 5(b).

 figure: Fig. 5.

Fig. 5. (a) Ring resonator under dynamic modulation by the electro-optic modulator (EOM), and in addition coupled to two external waveguides. (b) and (c) Detected signals in external waveguide 2: the top panels are the input spectra injected through external waveguide 1. The ring is modulated for a period from 0 to tf. The output spectra, after the modulation is turned off, are plotted in the remaining panels. In (b), the modulation frequency periodically switched between Δ=±0.055 c/n0L as described in Fig. 4(b). In (c), the modulation frequency is maintained constant at Δ=0.055 c/n0L.

Download Full Size | PDF

4. DISCUSSION AND CONCLUSION

We discuss a few experimental aspects for on-chip implementation of our theoretical predictions. Modulation-induced coupling between different resonant modes separated by a free spectral range was demonstrated in [31], which used a silicon ring resonator with a radius of 445 μm modulated at a frequency of Ω=26GHz. Our theory can be implemented in a similar system. The maximum coupling strength between the modes, as observed in Ref. [31], translates to α0.8 in Eq. (11) as applied to this system. Therefore, our choice of α=0.2 in the simulation is within what one can accomplish experimentally. To observe a period of Bloch oscillation, the photon lifetime τp in the ring needs to be larger than the period of a single Bloch oscillation 2π/Δ, which, with a choice of Δ0.1ΩR2GHz, translates into a requirement of the quality factor of the ring of the order of 105106. Such a quality factor may be achievable in silicon ring modulators [31]. Alternatively, one can consider other material systems that have significant electro-optic effects, and where high-Q resonators have already been achieved [36]. To demonstrate the Hamiltonian of Eq. (2), there needs to be sufficient bandwidth over which the group velocity dispersion is small. For a silicon waveguide, near-zero group velocity dispersion has been observed over a bandwidth of 670 nm [37], which is far larger than what we require here.

Our study in general points to interesting opportunities for the control of light spectra through the use of dynamic modulation inside the cavity. While the simulation above is for a classical wave, since the phase modulator as described by Eq. (8) is a linear and lossless device, the same physics should occur in the quantum regime at the few photon level, as well. The unidirectional shift here is, therefore, potentially useful for photon frequency translation [27]. Importantly, here the magnitude of frequency translation can be far larger than the modulation frequency. Our work here also indicates other interesting new opportunities for manipulating the spectrum of light in microring resonators. For example, with a different choice of parameters, an effect of dynamic localization [29,35,38,39] in the frequency domain may be achieved. Our work here points to interesting new opportunities for manipulating the spectrum of light in microring resonators.

Funding

Air Force Office of Scientific Research (AFOSR) (FA9550-12-1-0488).

REFERENCES

1. K. J. Vahala, “Optical microcavities,” Nature 424, 839–846 (2003). [CrossRef]  

2. Q. Xu, S. Sandhu, M. L. Povinelli, J. Shakya, S. Fan, and M. Lipson, “Experimental realization of an on-chip all-optical analogue to electromagnetically induced transparency,” Phys. Rev. Lett. 96, 123901 (2006). [CrossRef]  

3. A. B. Matsko, A. A. Savchenkov, W. Liang, V. S. Ilchenko, D. Seidel, and L. Maleki, “Collective emission and absorption in a linear resonator chain,” Opt. Express 17, 15210–15215 (2009). [CrossRef]  

4. T. Hu, P. Yu, C. Qiu, H. Qiu, F. Wang, M. Yang, X. Jiang, H. Yu, and J. Yang, “Tunable Fano resonances based on two-beam interference in microring resonator,” Appl. Phys. Lett. 102, 011112 (2013). [CrossRef]  

5. B. Peng, S. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys. 10, 394–398 (2014). [CrossRef]  

6. T. J. Kippenberg, R. Holzwarth, and S. A. Diddams, “Microresonator-based optical frequency combs,” Science 332, 555–559 (2011). [CrossRef]  

7. K. Saha, Y. Okawachi, J. S. Levy, R. K. W. Lau, K. Luke, M. A. Foster, M. Lipson, and A. L. Gaeta, “Broadband parametric frequency comb generation with a 1-μm pump source,” Opt. Express 20, 26935–26941 (2012). [CrossRef]  

8. S.-W. Huang, H. Zhou, J. Yang, J. F. McMillan, A. Matsko, M. Yu, D.-L. Kwong, L. Maleki, and C. W. Wong, “Mode-locked ultrashort pulse generation from on-chip normal dispersion microresonators,” Phys. Rev. Lett. 114, 053901 (2015). [CrossRef]  

9. A. R. Johnson, Y. Okawachi, M. R. E. Lamont, J. S. Levy, M. Lipson, and A. L. Gaeta, “Microresonator-based comb generation without an external laser source,” Opt. Express 22, 1394–1401 (2014). [CrossRef]  

10. S. B. Papp, K. Beha, P. Del’haye, F. Quinlan, H. Lee, K. J. Vahala, and S. A. Diddams, “Microresonator frequency comb optical clock,” Optica 1, 10–14 (2014). [CrossRef]  

11. A. G. Griffith, R. K. W. Lau, J. Cardenas, Y. Okawachi, A. Mohanty, R. Fain, Y. H. D. Lee, M. Yu, C. T. Phare, C. B. Poitras, A. L. Gaeta, and M. Lipson, “Silicon-chip mid-infrared frequency comb generation,” Nat. Commun. 6, 6299 (2015). [CrossRef]  

12. J. N. Winn, S. Fan, J. D. Joannopoulos, and E. P. Ippen, “Interband transitions in photonic crystals,” Phys. Rev. B 59, 1551–1554 (1999). [CrossRef]  

13. P. Dong, S. F. Preble, J. T. Robinson, S. Manipatruni, and M. Lipson, “Inducing photonic transitions between discrete modes in a silicon optical microcavity,” Phys. Rev. Lett. 100, 033904 (2008). [CrossRef]  

14. Z. Yu and S. Fan, “Complete optical isolation created by indirect interband photonic transitions,” Nat. Photonics 3, 91–94 (2009). [CrossRef]  

15. T. Ozawa, H. M. Price, N. Goldman, O. Zilberberg, and I. Carusotto, “Synthetic dimensions in integrated photonics: From optical isolation to four-dimensional quantum Hall physics,” Phys. Rev. A 93, 043827 (2016). [CrossRef]  

16. L. Yuan, Y. Shi, and S. Fan, “Photonic gauge potential in a system with a synthetic frequency dimension,” Opt. Lett. 41, 741–744 (2016). [CrossRef]  

17. G. Lenz, I. Talanina, and C. M. de Sterke, “Bloch oscillations in an array of curved optical waveguides,” Phys. Rev. Lett. 83, 963–966 (1999). [CrossRef]  

18. H. Trompeter, W. Krolikowski, D. N. Neshev, A. S. Desyatnikov, A. A. Sukhorukov, Y. S. Kivshar, T. Pertsch, U. Peschel, and F. Lederer, “Bloch oscillations and Zener tunneling in two-dimensional photonic lattices,” Phys. Rev. Lett. 96, 053903 (2006). [CrossRef]  

19. S. Longhi, “Optical Zener-Bloch oscillations in binary waveguide arrays,” Europhys. Lett. 76, 416–421 (2006). [CrossRef]  

20. F. Dreisow, A. Szameit, M. Heinrich, T. Pertsch, S. Nolte, A. Tünnermann, and S. Longhi, “Bloch-Zener oscillations in binary superlattices,” Phys. Rev. Lett. 102, 076802 (2009). [CrossRef]  

21. M. Levy and P. Kumar, “Nonreciprocal Bloch oscillations in magneto-optic waveguide arrays,” Opt. Lett. 35, 3147–3149 (2010). [CrossRef]  

22. U. Peschel, C. Bersch, and G. Onishchukov, “Discreteness in time,” Open Phys. 6, 619–627 (2008). [CrossRef]  

23. C. Bersch, G. Onishchukov, and U. Peschel, “Experimental observation of spectral Bloch oscillations,” Opt. Lett. 34, 2372–2374 (2009). [CrossRef]  

24. F. Bloch, “Über die Quantenmechanik der Elektronen in Kristallgittern,” Z. Phys. 52, 555–600 (1929). [CrossRef]  

25. E. Haller, R. Hart, M. J. Mark, J. G. Danzl, L. Reichsöllner, and H.-C. Nägerl, “Inducing transport in a dissipation-free lattice with super Bloch oscillations,” Phys. Rev. Lett. 104, 200403 (2010). [CrossRef]  

26. M. T. Rakher, L. Ma, O. Slattery, X. Tang, and K. Srinivasan, “Quantum transduction of telecommunications-band single photons from a quantum dot by frequency upconversion,” Nat. Photonics 4, 786–791 (2010). [CrossRef]  

27. H. J. McGuinness, M. G. Raymer, C. J. McKinstrie, and S. Radic, “Quantum frequency translation of single-photon states in a photonic crystal fiber,” Phys. Rev. Lett. 105, 093604 (2010). [CrossRef]  

28. P. Lodahl, S. Mahmoodian, and S. Stobbe, “Interfacing single photons and single quantum dots with photonic nanostructures,” Rev. Mod. Phys. 87, 347–400 (2015). [CrossRef]  

29. L. Yuan and S. Fan, “Three-dimensional dynamic localization of light from a time-dependent effective gauge field for photons,” Phys. Rev. Lett. 114, 243901 (2015). [CrossRef]  

30. N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders College, 1976).

31. L. D. Tzuang, M. Soltani, Y. H. D. Lee, and M. Lipson, “High RF carrier frequency modulation in silicon resonators by coupling adjacent free-spectral-range modes,” Opt. Lett. 39, 1799–1802 (2014). [CrossRef]  

32. H. A. Haus, Waves and Fields in Optoelectronics (Prentice-Hall, 1984).

33. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics, 2nd ed. (Wiley, 2007).

34. H. A. Haus, “Mode-locking of lasers,” IEEE J. Sel. Top. Quantum Electron. 6, 1173–1185 (2000). [CrossRef]  

35. D. H. Dunlap and V. M. Kenkre, “Dynamic localization of a charged particle moving under the influence of an electric field,” Phys. Rev. B 34, 3625–3633 (1986). [CrossRef]  

36. A. A. Savchenkov, A. B. Matsko, V. S. Ilchenko, I. Solomatine, D. Seidel, and L. Maleki, “Tunable optical frequency comb with a crystalline whispering gallery mode resonator,” Phys. Rev. Lett. 101, 093902 (2008). [CrossRef]  

37. L. Zhang, Q. Lin, Y. Yue, Y. Yan, R. G. Beausoleil, and A. E. Willner, “Silicon waveguide with four zero-dispersion wavelengths and its application in on-chip octave-spanning supercontinuum generation,” Opt. Express 20, 1685–1690 (2012). [CrossRef]  

38. S. Longhi, “Self-imaging and modulational instability in an array of periodically curved waveguides,” Opt. Lett. 30, 2137–2139 (2005). [CrossRef]  

39. A. Joushaghani, R. Iyer, J. K. S. Poon, J. S. Aitchison, C. M. de Sterke, J. Wan, and M. M. Dignam, “Generalized exact dynamic localization in curved coupled optical waveguide arrays,” Phys. Rev. Lett. 109, 103901 (2012). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) Ring resonator modulated by an electro-optic modulator (EOM). (b) In the absence of group velocity dispersion, the ring in (a) supports a set of resonant modes (blue dashed lines) with their resonant frequencies equally spaced by Ω R . The modulation frequency Ω is chosen to be slightly different from Ω R . For the modes that are located at ω 0 , the modulation generates sidebands that are detuned from the other resonant modes. (c) The system as described in (b) can be mapped onto a tight-binding model of a photon under an effective force [Eq. (4)].
Fig. 2.
Fig. 2. Evolution in time of the total energy for each resonant mode ( I m ( t ) ring | E m ( t , z ) | 2 d z ): (a)  Δ = 0.1 c / n 0 L , and (b)  Δ = 0.5 c / n 0 L , when only the m = 0 mode is excited.
Fig. 3.
Fig. 3. Evolution in time of the total energy for each resonant mode ( I m ( t ) ring | E m ( t , z ) | 2 d z ): (a)  Δ = 0.055 c / n 0 L , and (b)  Δ = 0.055 c / n 0 L , when several modes are initially excited.
Fig. 4.
Fig. 4. (a) The evolution in time of the total energy for each resonant mode ( I m ( t ) ) , as the modulation frequency is periodically switched between Δ = ± 0.055 c / n 0 L , as described in (b).
Fig. 5.
Fig. 5. (a) Ring resonator under dynamic modulation by the electro-optic modulator (EOM), and in addition coupled to two external waveguides. (b) and (c) Detected signals in external waveguide 2: the top panels are the input spectra injected through external waveguide 1. The ring is modulated for a period from 0 to t f . The output spectra, after the modulation is turned off, are plotted in the remaining panels. In (b), the modulation frequency periodically switched between Δ = ± 0.055 c / n 0 L as described in Fig. 4(b). In (c), the modulation frequency is maintained constant at Δ = 0.055 c / n 0 L .

Equations (16)

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

ω m = ω 0 + m Ω R ,
H = m ω m a m a m + m 2 g cos ( Ω t ) ( a m a m + 1 + a m + 1 a m ) ,
H RWA = g m ( e i Δ t c m c m + 1 + e i Δ t c m + 1 c m ) ,
H ˜ RWA = g m ( c m c m + 1 + c m + 1 c m ) + m m Δ c m c m ,
| Ψ = m v m c m | 0 | Ψ ˜ = m v ˜ m c m | 0 = m v m e i m Δ t c m | 0 ,
E ( t , r , z ) = m E m ( t , z ) E m ( r ) e i ω m t ,
( z + i β ( ω m ) ) E m ( t , z ) n g ( ω m ) c t E m ( t , z ) = 0 ,
E ( t + , r , z 0 ) = E ( t , r , z 0 ) e i α cos ( Ω t ) ,
E m ( t + , z 0 ) = J 0 ( α ) E m ( t , z 0 ) + q i q J q ( α ) [ E m + q ( t , z 0 ) e i q Δ t + E m q ( t , z 0 ) e i q Δ t ] ,
Δ E m ( t , z 0 ) = i α 2 [ E m + 1 ( t , z 0 ) e i Δ t + E m 1 ( t , z 0 ) e i Δ t ] .
T R E m T = i α 2 ( E m + 1 e i Δ T + E m 1 e i Δ T ) .
E m ( t + , z s ) = E m ( t , z s ) + S m ( t ) .
S m ( t ) = f m e 2 ln 2 ( t 20 ) 2 / 100 ,
f m = exp [ ( ω m ω c ) 2 / ( 5 Ω R ) 2 ] ,
E m ( t + , z 1,2 ) = 1 γ 1,2 2 ( t ) E m ( t , z 1,2 ) i γ 1,2 ( t ) E m ( 1,2 ) ( t , z 1,2 ) ,
E m ( 1,2 ) ( t + , z 1,2 ) = 1 γ 1,2 2 ( t ) E m ( 1,2 ) ( t , z 1,2 ) i γ 1,2 ( t ) E m ( t , z 1,2 ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.