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

Raman-Kerr frequency combs in microresonators with normal dispersion

Open Access Open Access

Abstract

We generalize the coupled mode formalism to study the generation of frequency combs in microresonators with simultaneous Raman and Kerr nonlinearities and investigate an impact of the former on the formation of frequency combs and dynamics of platicons in the regime of the normal group velocity dispersion. We demonstrate that the Raman effect initiates generation of sidebands, which cascade further in four-wave mixing and reshape into the Raman-Kerr frequency combs. We reveal that the Raman scattering induces a strong instability of the platicon pulses associated with the Kerr effect and normal dispersion. This instability results in branching of platicons and complex spatiotemporal dynamics.

© 2017 Optical Society of America

1. Introduction

High-Q optical microresonators opened a new pathway to the development of Kerr frequency comb generators with characteristics unattainable in the comb systems based on mode-locked lasers [1]. Microresonator combs are generated by coupling a CW laser light into a resonator that converts the pump frequency into a broadband comb through the cascaded four-wave mixing (FWM) due to Kerr nonlinearity of the medium boosted by the strong light confinement [2]. The spacing between the comb lines corresponding to the free spectral range (FSR) of the resonator is determined by the inverse of the resonator round-trip time and can range from few GHz to THz [3]. The threshold pump power which can be below 1 mW level [4] is determined by the resonator quality-factor Q and mode volume. The microresonator group velocity dispersion (GVD) plays a major role in the comb generation processes since it determines the bandwidth of the four-wave mixing gain [5]. In particular, microresonators with the anomalous GVD have been used to demonstrate octave spanning combs [6,7]. Finally, highly-coherent mode-locked combs based on bright dissipative Kerr solitons [8] promise a plethora of applications. However, most of the dielectric materials used for microresonators have normal GVD in the visible and mid-IR spectral ranges, prohibiting four-wave mixing gain and hence coherent and broadband soliton combs. Comb generation in the normal GVD range, therefore, is governed by other physical mechanisms and it has also been explored actively in recent years [9,10]. In particular, local dispersion perturbation due to linear coupling between various families of the resonators’ modes [11–13] allows excitation of a new type of dissipative soliton structures, the so-called platicons [14–16], associated with normal GVD combs. Platicons represent the bound states of the opposing switching waves [17] connecting the upper and lower states of the bistability loop and satisfying periodic boundary conditions in a microresonator. Interestingly, quite similar structures were discovered for complex Ginzburg-Landau equation, where bound kink–antikink complexes can also form top-flat pulse [18]. Properties of platicons still have not been explored to the same degree of details as the physics of the bright comb solitons. In particular, the influence of the Raman effect on the normal GVD combs and platicons remains largely unknown.

The interplay between the Kerr and Raman effects in microresonators has attracted significant interest in recent years, especially in the context of bright comb solitons in the anomalous GVD regime [19–25]. The above works reported the soliton self-frequency shift [19, 22–24], interplay of the Cherenkov radiation and Raman effects [19] and the generation of the two-color Stokes solitons [25]. Most of these studies are based on the generalized Lugiato-Lefever equation (LLE) [26] incorporating an approximation of the Raman gain by the first derivative of the pulse intensity in the physical space [27–29], corresponding to an unrealistic linear increase of the gain with frequency detuning. In this paper, we derive a model from the first principles, describing optomechanical coupling with Raman mechanical oscillator [19,30] and proceed to solve it with the coupled-mode approach [31], that incorporates both Kerr and Raman nonlinearities. The resulting equations are used to investigate comb generation within the gain line of the Raman scattering and the associated dynamics of platicons. Our present approach appears to be useful for combs with a small number of modes and complex dispersions and also for the interpretation of the interplay between Raman and Kerr effects. We demonstrate that a significant excess of the Raman gain over losses leads to instabilities of platicons appearing as a branching of the parent platicon into a bifurcation tree with newly generated stable and unstable platicons. The associated comb generation appears to be qualitatively different from the one known for bright solitons where the Raman effect mostly results in self-frequency shift [22,23].

2. Theory

A part of the total susceptibility χ of the microresonator’s material, which is induced by the Raman effect, is described by the equation for a classical mechanical oscillator with generalized amplitude q driven by the light intensity, see, e.g., [19, 30]: χ = χ(1) + χ(3)|E|2, where χ(1) includes the linear susceptibility χ0(1) and the amplitude q of the Raman oscillator:

χ(1)=χ0(1)+χ(1)qq,
d2qdt2+2γdqdt+Ω02q=wFq,
where w is the inverse of the oscillator mass, Ω0 corresponds to a Stokes resonance with a width of 2γ at half-maximum [see Fig. 1], and generalized force Fq corresponding to optical field:
Fq=02χ(1)q|E|2Veff,
where 0 is the vacuum permittivity and Veff is the mode volume.

 figure: Fig. 1

Fig. 1 A schematic representation of Raman scattering in microresonators. The Raman peak covers several free spectral ranges (FSR) in a large diameter microresonator. ΩR is the detuning of the Raman peak from the pump frequency (ΩR2π=9.66THzfor CaF2 and 12.3 THz for MgF2).

Download Full Size | PDF

For the optical part we start from the nonlinear wave equation [31]:

××E+n2c22Et2=10c22Ppt21c2χ(1)q2t2qEχ(3)c22(E|E|2)t2,
where Pp=0χ(fp(r)eiω0t) is the polarization induced by the pump field. We expand the electric field E⃗ in terms of the resonator modes e⃗μ(r):
E(r,t)=(eiωtμAμ(t)eμ(r)),
where we accounted detunings of the mode eigenfrequencies ωμ from the pump frequency ω in time dependent amplitudes Aμ(t) expecting that all modes are oscillating with the pump frequency. |μ| = 0, 1, 2, . . . is the mode numbers, where μ = 0 corresponds to the mode closest to the pump. We assume, that the sum goes from μ = −N to N, with unspecified summation limits. The mode profiles are defined as
××eμn2ωμ2c2eμ=0
with the orthogonality relation eμ*eν=δμν. The Raman force Fq is expressed in terms of the modal amplitudes as:
Fq=04χ(1)qμν(eμeνAμAνe2iωt+eμeν*AμAν*)Veff.
Neglecting fast-oscillating terms in Eq. (7) and using q=weν(r)eν*(r)qνν(t) we transform the equation for the mechanical oscillator (2) to the form
d2qννdt2+2γdqννdt+Ω02qνν=w04Veffχ(1)qAνAν*.
We can think of qνν′ as an amplitude of oscillations of the Raman oscillator excited by a force with a form of the dot product of ν and ν′ modes. After tedious, but straightforward algebra, which includes the standard omission of the second derivatives of the slow varying amplitudes and of the fast oscillating exponential terms, we derive the system of equations for the modal amplitudes coupled to the Raman oscillator amplitudes:
A˙μ=iωμ2ω22ωAμ+iωfμχ(1)2n2δμ0+i3χ(3)ω8n2μμμΛμμμμAμAμAμ*+iw2n2χ(1)qωννμΛμμννqνν(t)Aμ.
Here Λμμνν=(eνeν*)(eμeμ*)dV. We then can multiply Eq. (8) by Λμμνν and sum over ν and ν′ to simplify the Raman term sum in (9). To further simplify Λμμνν we should have some insight into mode forms and structure. We presume that we get nonzero Λ ≈ 1 only for ν′ = ν + μ′μ. This relation is supposed to be correct for all cavities, which modes can be represented by complex exponent in one direction (standing or traveling waves) and fixed field distribution in the others. For illustration we consider whispering gallery modes (WGMs). First, using the cylindrical symmetry of the WGM problem we extract the azimuthal dependence as e⃗ν(r⃗) → e⃗p,q,ν(r, z)eiνφ. Here p and q are the transverse mode numbers that were previously implicit in ν and μ, while ν and μ are now pure azimuthal indexes. If ν is very large (which is usually the case), then the transverse profiles of all the modes can be assumed to be similar and independent of ν. Performing azimuthal integration we get nonzero Λ ≈ 1 only for ν′ = ν + μ′μ and the same p and q of the four modes. Note that due to the latter the following applies only to each single mode family separately. Renormalizing (A2ω0VeffA) and introducing the loss terms κμ into Eq. (9) we obtain a system of equations for modal amplitudes
Aμt=i(ωμω)Aμκμ2Aμ+iGωμqμμ(t)Aμ+igμ=μ+μμAμAμAμ*+Fδμ0
and Raman oscillations amplitudes
d2qμμdt2+2γdqμμdt+Ω02qμμ=GωνAνAν+μμ*.
Here, g=ω02cn2n2Veff is the nonlinear coupling coefficient, n2 is the nonlinear refractive index, Veff is the effective mode volume, κ=ω0Q=κ0+κc denotes the cavity decay rate as the sum of intrinsic decay rate κ0 and coupling rate κc, Q is the total quality factor, F=κcP0ω is the pump amplitude, G=w2n2χ(1)q is the Raman gain coefficient. Now, Eq. (10) coincides with the model used in [8] with the additional Raman term.

Expanding the eigenfrequencies ωμ=ω0+μD1+μ2D22 and introducing the field envelope A=μAμeiμϕ, where ϕ is the polar angle, we find that Eq. (10) transforms to the Lugiato-Lefever equation:

At=i[(ω0ω)iκ2]AD1Aϕ+iD222Aϕ2+igA|A|2+iGqωA+F,
where qμμAμ is replaced by q(ϕ)A(ϕ), where q(ϕ)=qηeiηϕ. The set of Raman equations (11) is replaced under these conditions with
d2q(ϕ,t)dt2+2γdq(ϕ,t)dt+Ω02q(ϕ,t)=Gω|A(ϕ,t)|2.
Therefore, the modal approach presented above is equivalent to the system of the Lugiato-Lefever and the Raman oscillator equations. A formal solution of Eq. (13) is given by the convolution of its right hand side with the oscillator’s response function, h(t):q(ϕ,t)=+h(tt)|A(ϕ,t)|2H(tt)dt, where h(tt)=Geγ(tt)sinΩ02γ2(tt) and H (tt′) is Heaviside step function. A linear in frequency approximation for h(Ω)=+h(t)eiΩtdtG(1ΩτR) leads to the Raman shock term −iG2ħωτRA∂t|A|2 in the time domain. This approach may be suitable to study the wave forms with the spectra centered away from the maximum of the Raman gain, see, e.g., [23], but can not be applied for the type of dynamics reported below, which is initiated by the sidebands growing at the frequencies determined by the maximum of the gain and by the width of the Raman line.

For the numerical simulation Eq. (10) is converted to a dimensionless form through substitutions Aμ=D1/gaμeiμD1t and τ = tD1, so that modes are now expected to be separated by FSR, and time is dimensionless:

aμτ=(κ2D1+iζμ)aμ+Rμ+iμμ(2δμμ)aμaμaμ+μμ*+fδμ0,
All mode numbers μ are defined relative to the pump mode μ = mm0 with the initial azimuthal number m02πrn0λ, where λ=2πcω0. The cold cavity dispersion law is ωμ=ω0+kDkk!μk:D1cn0R is the FSR (R is the radius of the resonator) and D2 < 0 (normal GVD), ζμ=ωμωμD1D1 is the normalized detuning, f=gηκP0D13ω is the dimensionless pump amplitude and Rμ=iωGD1μqμμaμei(μμ)D1t is the Raman term.

To this end, we rewrite Eq. (11) for the Raman oscillator. First, we note that due to the resonant character of the equation we can reduce the summation only to those terms that lie inside the resonance band |μμ′|D1 ∈ Ω0 ± 2γ. Second, as Eq. (14) has constant parameters, only slowly-varying parts of Raman term will influence the solution. Hence one does not need to solve separate equations in (11) for each η = μμ′, and we can use the single equation obtained after the summation over η:

d2q¯dτ2+2γ˜dq¯dτ+Ω˜02q¯=ηD1Ω0+2γ,νaνaνη*eiητ,
where q¯=gD1Gωqη. Here we also performed the same renormalization of time as in (10): γ˜=γD1 and Ω˜0=ΩD1 and made dimensionless. So the Raman term in Eq. (14) becomes
Rμ=iGRq¯μaμeiητ,
where GR=G2ω2D12g=wVeff4n2cn2D12(χ(1)q)2 is the normalized Raman gain coefficient.

3. Numerical simulation

Our numerical model is based on a system of dimensionless coupled mode equations (14) and the oscillator equation (15) with the Raman term (16). In our simulations, we used at least 1024 modes and solved the system using the adaptive Runge-Kutta integrator. The integration time was large enough for all transient effects to disappear. Nonlinear terms were calculated using a fast Fourier method proposed in [32]. Parameters used below correspond to a crystalline microresonator [14–16]: f = 0.0014, κ2D1=0.005, D2κ=0.01. For the frequency and damping rate of the Raman oscillator we assumed Ω̃0 = 200 and γ̃ = 1.5. These parameters do not correspond to a specific material, but are realistic for crystalline materials. However qualitatively the same results were observed for other parameters of the Raman oscillator.

We studied the comb generation assuming normal GVD and applying the pump detuning corresponding to the upper branch of the steady-state intracavity field perturbed by noise. The Raman gain was used as a control parameter. For a given pump frequency providing that the Raman gain GR is large enough we observed combs with a spacing defined by the Raman frequency. At the initial stage of evolution, we observed the generation of primary Stokes and anti-Stokes lines at μs = −198 and μs = 202, which then evolved due to a cascaded generation of secondary lines [see Fig. 2]. We note that the position of the Raman peaks does not exactly match Ω̃ = 200 due to non-zero GVD responsible for the non-equidistance of the spectrum. A mode with μ = 198 for the D2κ=0.01 approximately corresponds to 200D1. This distinguishes this state from a soliton comb, where the distance between the lines is accurately determined by FSR due to mode locking. One can also see that the Stokes part of the generated comb is stronger than the anti-Stokes one (Fig. 2). We have checked numerically that the Kerr part of the four-wave mixing has weak effect on the generation of the frequency comb. However, the fact that anti-Stokes lines are non-negligible strongly suggests that the comb generation was also influenced by the Raman part of FWM process. The Kerr nonlinearity manifests itself mainly in self-action effects.

 figure: Fig. 2

Fig. 2 Optical spectrum for the primary Stokes line formation for GR = 120 (a) and GR = 160 (b). The Raman peak overlaps the mode with μs = −198, the width of the Raman gain γ̃ = 1.5, ζ = 0.02. The anti-Stokes line is strongly suppressed. The Raman FWM between the pump mode and the Stokes mode creates an anti-Stokes peak at the mode number μas = 202 and cascades a peak with frequency such as ωcascade is divisible by the Raman frequency. Spectral lines around the Raman peak are shown in the insets.

Download Full Size | PDF

Lines around the pumped mode line result from nondegenerate FWM interactions between the primary Stokes and anti-Stokes components [see Fig. 3]. The number of the generated comb lines depends on the Raman line-width γ̃ and gain GR. The increase in the Raman gain value leads to the growth of the lines number. The detuning range providing comb generation also becomes wider with the increase of GR. It is also worth noting that the Raman line-width in a very small microresonator (diamond or silicon) can be much narrower than the FSR. If the peak of the Raman gain is far detuned from any cavity resonance, then the Raman effect will be either strongly suppressed or practically non-existent [33]. Contrary, in large resonators and for materials with wider Raman lines, the Raman gain can cover many FSRs. For example, Raman line overlaps tens of FSRs for crystalline microresonators (full-width at half maximum line-width γ2π=450GHz for calcium fluoride, γ2π=210GHz for magnesium fluoride).

 figure: Fig. 3

Fig. 3 The comb spectrum is generated by a combination of the Raman scattering, which downshifts its frequency and nondegenerate four-wave mixing processes in which all four photons have different frequencies. Insets show energy level diagrams for the Raman scattering and four-wave mixing processes. The Raman part of the FWM plays a dominant role in the spectrum formation.

Download Full Size | PDF

The generation of a Raman comb is also possible if a pump frequency scan is applied. This approach can be more efficient since in this case, the system can stay in the upper state for larger detunings, which corresponds to larger intensities of the pumped mode providing more efficient Raman process. In contrast, starting from the noise-like inputs at large values of detuning one may generate only low-intensity steady-state solutions not always supporting efficient generation of cascaded Raman peaks.

4. Platicon dynamics

In this section, we study the generation of platicons in the presence of Raman effect. We do not expect that the Raman effect on its own can lead to the platicon generation since platicons are known to require an application of either the pumped mode shift method [14] or the pump modulation [15]. Before looking at non-trivial effects in normal dispersion regime, we first simulated the process of bright soliton propagation in the anomalous regime for a range of the Raman gain values, while fixing Ω̃0 = 100 and γ̃ = 1.0. As it was demonstrated (e.g. in [19]) the presence of the Raman gain leads to the modification of the soliton profile and to the appearance of nonzero “transverse” (drift) velocity. It was found that the drift velocity increases with the growth of Raman gain coefficient GR. Such a drift may be explained by the fact that low-frequency components of the soliton spectrum may experience significant Raman gain. Our simulations in anomalous regime were in agreement with known results, demonstrating stable propagation of bright solitons with different velocities and frequency shifts corresponding to selected values of the Raman gain.

To study the platicon dynamics for the same set of parameters, we modified the conventional dispersion law as ωμ=ω0+δ0μΔ+D1μ+D22μ2 by including the pumped mode shift Δ = 2κ and taking D2 < 0. Our numerical simulations revealed that the Raman effect strongly influences the platicon dynamics in a way very different from that of bright solitons [see Fig. 4]. In particular, the velocity and frequency shifts typical for bright solitons do not appear for platicons. Platicons seem to be insensitive to the Raman process up to some threshold value of the Raman gain [see Fig. 4(d)]. Note, that the third order dispersion, unlike the Raman effects, acts on platicons [16] in a way qualitatively similar to its action on the bright solitons [19] and leads to a non-zero drift velocity.

 figure: Fig. 4

Fig. 4 Wide platicon (ζ0 = 0.02) evolution for different values of the Raman gain: (a) GR = 150; (b) GR = 175; (c) GR = 200; (d) GR = 220; (e) GR = 300; (f) GR = 575.

Download Full Size | PDF

If GR exceeds some critical value, then platicons lose their stability and give birth to various spatiotemporal patterns [Figs. 4(b)–4(d)]. As it was shown in [14] the width of the platicon spectrum varies with pump frequency detuning. Using different values of ζ0 we studied the dynamics of different platicons in Raman scattering band. Numerical simulations demonstrate that the Raman induced instability of the relatively wide platicons typically leads to splitting (branching) [Figs. 4(b)–4(d), Fig. 5], while the narrow ones decay during propagation [Figs. 5(d)–5(f)]. In some cases a kind of quasi-stable two-hump solutions was observed as seen in Fig. 4(d).

 figure: Fig. 5

Fig. 5 Platicon dynamics for intermediate width (ζ0 = 0.03, a–c) and narrow (ζ0 = 0.04, d–f) platicon with different values of the Raman gain: (a) GR = 200; (b) GR = 250; (c) GR = 300, (d) GR = 400; (e) GR = 575; (f) GR = 700.

Download Full Size | PDF

Further growth of the Raman gain leads to a blurring of high-intensity patterns, [Fig. 4(e)], and to a transformation of a platicon into a complex spatiotemporal pattern [Fig. 4(f)]. Obviously, the greater the Raman gain is, the faster the instability develops. The instability also leads to a significant distortion of the platicon spectrum. For parameters used at Fig. 4(b) one can compare initial platicon spectrum [Fig. 6(a)] and the spectrum in the point of instability at τ = 5000 [Fig. 6(b)] which demonstrates the appearance of cascaded Raman peaks. The threshold of the platicon instability increases with the growth of the detuning and decrease of platicon width [compare Fig. 4(c) and Fig. 5(a), and Fig. 5(d)]. Also, for the same Raman gain, narrower platicons produce more structured intensity distribution profiles than the wider ones [compare Fig. 4(e) and Fig. 5(c)]. Since platicons are formed by binding of two counter-propagating fronts, [17], [34], it is natural to expect that the instabilities discussed here result from the Raman effect influence on the relative velocities of these fronts frustrating robustness of platicons.

 figure: Fig. 6

Fig. 6 Degradation of the platicon optical spectra for ζ0 = 0.02, GR = 175: (a) τ = 0; (b) τ = 5000. The Stokes mode corresponds to μ ≈ 100. The waveforms for the corresponding spectra are shown in the (c) and (d).

Download Full Size | PDF

The situation dramatically changes at some critical value of the platicon width (determined by pump frequency detuning). Under the influence of the high Raman gain exceeding some critical value, narrow platicons do not split into several parts and do not form complex structures during the propagation. For narrow platicons, the instability manifests itself as a decay of the initial profile to the low-intensity background, with the stronger Raman gain leading to the shorter platicon lifetimes [see Figs. 5(e)–5(f)].

5. Conclusion

We have developed the coupled mode approach to the Kerr frequency comb generation accounting for the Raman scattering in optical microresonators. Using this model, we demonstrated the generation of Kerr-Raman comb in the normal GVD regime and studied the platicon dynamics featuring branching instabilities leading to the formation of complex spatiotemporal structures.

Funding

Russian Science Foundation (17-12-01413).

Acknowledgments

D.V.S. acknowledges support from the ITMO University Visiting Professorship Scheme via the Government of Russian Federation Grant 074-U01. D.V.S. and M.L.G. acknowledge support from EU H2020 (691011, Soliring) for exchange visits. A.V.C. acknowledges scholarship from the Foundation for the Advancement of Theoretical Physics “BASIS”.

References and links

1. P. Del’Haye, A. Schliesser, O. Arcizet, T. Wilken, R. Holzwarth, and T. J. Kippenberg, “Optical frequency comb generation from a monolithic microresonator,” Nature 450, 1214–1217 (2007). [CrossRef]  

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

3. M. H. P. Pfeiffer, C. Herkommer, J. Q. Liu, H. R. Guo, M. Karpov, E. Lucas, M. Zervas, and T. J. Kippenberg, “Octave-spanning dissipative Kerr soliton frequency combs in Si3N4 microresonators,” Optica 4, 684–691 (2017). [CrossRef]  

4. A. A. Savchenkov, A. B. Matsko, D. Strekalov, M. Mohageg, V. S. Ilchenko, and L. Maleki, “Low threshold optical oscillations in a whispering gallery mode CaF2 resonator,” Phys. Rev. Lett. 93, 243905 (2004). [CrossRef]  

5. T. Herr, V. Brasch, J. D. Jost, I. Mirgorodskiy, G. Lihachev, M. L. Gorodetsky, and T. J. Kippenberg, “Mode spectrum and temporal soliton formation in optical microresonators,” Phys. Rev. Lett. 113, 123901 (2014). [CrossRef]   [PubMed]  

6. Y. Okawachi, K. Saha, J. S. Levy, Y. H. Wen, M. Lipson, and A. L. Gaeta, “Octave-spanning frequency comb generation in a silicon nitride chip,” Opt. Lett. 36, 3398–3400 (2011). [CrossRef]   [PubMed]  

7. P. Del’Haye, T. Herr, E. Gavartin, M. L. Gorodetsky, R. Holzwarth, and T. J. Kippenberg, “Octave spanning tunable frequency comb from a microresonator,” Phys. Rev. Lett. 107, 063901 (2011). [CrossRef]  

8. T. Herr, V. Brasch, J. D. Jost, C. Y. Wang, N. M. Kondratiev, M. L. Gorodetsky, and T. J. Kippenberg, “Temporal solitons in optical microresonators,” Nat. Photon. 8, 145–152 (2014). [CrossRef]  

9. X. Xue, Y. Xuan, P.-H. Wang, Y. Liu, D. E. Leaird, M. Qi, and A. M. Weiner, “Normal-dispersion microcombs enabled by controllable mode interactions,” Las. Photon. Rev. 9, L23–L28 (2015).

10. X. Xue, M. Qi, and A. M. Weiner, “Normal-dispersion microresonator Kerr frequency combs,” Nanophotonics 5, 244–262 (2016). [CrossRef]  

11. A. A. Savchenkov, A. B. Matsko, W. Liang, V. S. Ilchenko, D. Seidel, and L. Maleki, “Kerr frequency comb generation in overmoded resonators,” Opt. Express 20, 27290–27298 (2012). [CrossRef]   [PubMed]  

12. Y. Liu, Y. Xuan, X. Xue, P.-H. Wang, S. Chen, A. J. Metcalf, J. Wang, D. E. Leaird, M. Qi, and A. M. Weiner, “Investigation of mode coupling in normal-dispersion silicon nitride microresonators for Kerr frequency comb generation,” Optica 1, 137–144 (2014). [CrossRef]  

13. X. Xue, Y. Xuan, Y. Liu, P.-H. Wang, S. Chen, J. Wang, D. E. Leaird, M. Qi, and A. M. Weiner, “Mode-locked dark pulse Kerr combs in normal-dispersion microresonators,” Nat. Photon. 9, 594–600 (2015). [CrossRef]  

14. V. Lobanov, G. Lihachev, T. J. Kippenberg, and M. Gorodetsky, “Frequency combs and platicons in optical microresonators with normal GVD,” Opt. Express 23, 7713–7721 (2015). [CrossRef]   [PubMed]  

15. V. E. Lobanov, G. Lihachev, and M. L. Gorodetsky, “Generation of platicons and frequency combs in optical microresonators with normal GVD by modulated pump,” Europhysics Letters 112, 54008 (2015). [CrossRef]  

16. V. E. Lobanov, A. V. Cherenkov, A. E. Shitikov, I. A. Bilenko, and M. L. Gorodetsky, “Dynamics of platicons due to third-order dispersion,” Eur. Phys. J. D 71, 1–5 (2017). [CrossRef]  

17. P. Parra-Rivas, D. Gomila, E. Knobloch, S. Coen, and L. Gelens, “Origin and stability of dark pulse Kerr combs in normal dispersion resonators,” Opt. Lett. 41, 2402–2405 (2016). [CrossRef]   [PubMed]  

18. B. A. Malomed and A. A. Nepomnyashchy, “Kinks and solitons in the generalized Ginzburg-Landau equation,”, Phys. Rev. A 42, 6009–6014 (1990). [CrossRef]   [PubMed]  

19. C. Milián, A. V. Gorbach, M. Taki, A. V. Yulin, and D. V. Skryabin, “Solitons and frequency combs in silica microring resonators: Interplay of the Raman and higher-order dispersion effects,” Phys. Rev. A 92, 033851 (2015). [CrossRef]  

20. Y. K. Chembo, I. S. Grudinin, and N. Yu, “Spatiotemporal dynamics of Kerr-Raman optical frequency combs,” Phys. Rev. A 92, 043818 (2015). [CrossRef]  

21. G. Lin, S. Diallo, J. M. Dudley, and Y. K. Chembo, “Universal nonlinear scattering in ultra-high Q whispering gallery-mode resonators,” Opt. Express 24, 14880–14894 (2016). [CrossRef]   [PubMed]  

22. M. Karpov, H. Guo, A. Kordts, V. Brasch, M. H. P. Pfeiffer, M. Zervas, M. Geiselmann, and T. J. Kippenberg, “Raman self-frequency shift of dissipative Kerr solitons in an optical microresonator,” Phys. Rev. Lett. 116, 103902 (2016). [CrossRef]   [PubMed]  

23. X. Yi, Q.-F. Yang, K. Y. Yang, and K. Vahala, “Theory and measurement of the soliton self-frequency shift and efficiency in optical microcavities,” Opt. Lett. 41, 3419–3422 (2016). [CrossRef]   [PubMed]  

24. C. Bao, Y. Xuan, C. Wang, J. A. Jaramillo-Villegas, D. E. Leaird, M. Qi, and A. M. Weiner, “Soliton repetition rate in a silicon-nitride microresonator,” Opt. Lett. 42, 759–762 (2017). [CrossRef]   [PubMed]  

25. Q.-F. Yang, X. Yi, K. Y. Yang, and K. Vahala, “Stokes solitons in optical microcavities,” Nat. Phys. 13(1), 3875 (2016). [CrossRef]  

26. L. A. Lugiato and R. Lefever, “Spatial dissipative structures in passive optical systems,” Phys. Rev. Lett. 58, 2209–2211 (1987). [CrossRef]   [PubMed]  

27. Y. K. Chembo and C. R. Menyuk, “Spatiotemporal Lugiato-Lefever formalism for Kerr-comb generation in whispering-gallery-mode resonators,” Phys. Rev. A 87, 053852 (2013). [CrossRef]  

28. S. Coen, H. G. Randle, T. Sylvestre, and M. Erkintalo, “Modeling of octave-spanning Kerr frequency combs using a generalized mean-field Lugiato–Lefever model,” Opt. Lett. 38, 37–39 (2013). [CrossRef]   [PubMed]  

29. T. Hansson, D. Modotto, and S. Wabnitz, “Mid-infrared soliton and Raman frequency comb generation in silicon microrings,” Opt. Lett. 39, 6747–6750 (2014). [CrossRef]   [PubMed]  

30. A. V. Yulin and D. V. Skryabin, “Slowing down of solitons by intrapulse Raman scattering in fibers with frequency cutoff,” Opt. Lett. 31, 3092–3094 (2006). [CrossRef]   [PubMed]  

31. Y. K. Chembo and N. Yu, “Modal expansion approach to optical-frequency-comb generation with monolithic whispering-gallery-mode resonators,” Phys. Rev. A 82, 033801 (2010). [CrossRef]  

32. T. Hansson, D. Modotto, and S. Wabnitz, “On the numerical simulation of Kerr frequency combs using coupled mode equations,” Opt. Comm. 312, 134–136 (2014). [CrossRef]  

33. Y. Okawachi, M. Yu, V. Venkataraman, P. M. Latawiec, A. G. Griffith, M. Lipson, M. Lončar, and A. L. Gaeta, “Competition between Raman and Kerr effects in microresonator comb generation,” Opt. Lett. 42, 2786–2789 (2017). [CrossRef]   [PubMed]  

34. P. Parra-Rivas, E. Knobloch, D. Gomila, and L. Gelens, “Dark solitons in the Lugiato-Lefever equation with normal dispersion,” Phys. Rev. A 93, 063839 (2016). [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 (6)

Fig. 1
Fig. 1 A schematic representation of Raman scattering in microresonators. The Raman peak covers several free spectral ranges (FSR) in a large diameter microresonator. ΩR is the detuning of the Raman peak from the pump frequency ( Ω R 2 π = 9.66 THzfor CaF2 and 12.3 THz for MgF2).
Fig. 2
Fig. 2 Optical spectrum for the primary Stokes line formation for GR = 120 (a) and GR = 160 (b). The Raman peak overlaps the mode with μs = −198, the width of the Raman gain γ̃ = 1.5, ζ = 0.02. The anti-Stokes line is strongly suppressed. The Raman FWM between the pump mode and the Stokes mode creates an anti-Stokes peak at the mode number μas = 202 and cascades a peak with frequency such as ωcascade is divisible by the Raman frequency. Spectral lines around the Raman peak are shown in the insets.
Fig. 3
Fig. 3 The comb spectrum is generated by a combination of the Raman scattering, which downshifts its frequency and nondegenerate four-wave mixing processes in which all four photons have different frequencies. Insets show energy level diagrams for the Raman scattering and four-wave mixing processes. The Raman part of the FWM plays a dominant role in the spectrum formation.
Fig. 4
Fig. 4 Wide platicon (ζ0 = 0.02) evolution for different values of the Raman gain: (a) GR = 150; (b) GR = 175; (c) GR = 200; (d) GR = 220; (e) GR = 300; (f) GR = 575.
Fig. 5
Fig. 5 Platicon dynamics for intermediate width (ζ0 = 0.03, a–c) and narrow (ζ0 = 0.04, d–f) platicon with different values of the Raman gain: (a) GR = 200; (b) GR = 250; (c) GR = 300, (d) GR = 400; (e) GR = 575; (f) GR = 700.
Fig. 6
Fig. 6 Degradation of the platicon optical spectra for ζ0 = 0.02, GR = 175: (a) τ = 0; (b) τ = 5000. The Stokes mode corresponds to μ ≈ 100. The waveforms for the corresponding spectra are shown in the (c) and (d).

Equations (16)

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

χ ( 1 ) = χ 0 ( 1 ) + χ ( 1 ) q q ,
d 2 q d t 2 + 2 γ d q d t + Ω 0 2 q = w F q ,
F q = 0 2 χ ( 1 ) q | E | 2 V eff ,
× × E + n 2 c 2 2 E t 2 = 1 0 c 2 2 P p t 2 1 c 2 χ ( 1 ) q 2 t 2 q E χ ( 3 ) c 2 2 ( E | E | 2 ) t 2 ,
E ( r , t ) = ( e i ω t μ A μ ( t ) e μ ( r ) ) ,
× × e μ n 2 ω μ 2 c 2 e μ = 0
F q = 0 4 χ ( 1 ) q μ ν ( e μ e ν A μ A ν e 2 i ω t + e μ e ν * A μ A ν * ) V eff .
d 2 q ν ν d t 2 + 2 γ d q ν ν d t + Ω 0 2 q ν ν = w 0 4 V eff χ ( 1 ) q A ν A ν * .
A ˙ μ = i ω μ 2 ω 2 2 ω A μ + i ω f μ χ ( 1 ) 2 n 2 δ μ 0 + i 3 χ ( 3 ) ω 8 n 2 μ μ μ Λ μ μ μ μ A μ A μ A μ * + i w 2 n 2 χ ( 1 ) q ω ν ν μ Λ μ μ ν ν q ν ν ( t ) A μ .
A μ t = i ( ω μ ω ) A μ κ μ 2 A μ + i G ω μ q μ μ ( t ) A μ + i g μ = μ + μ μ A μ A μ A μ * + F δ μ 0
d 2 q μ μ d t 2 + 2 γ d q μ μ d t + Ω 0 2 q μ μ = G ω ν A ν A ν + μ μ * .
A t = i [ ( ω 0 ω ) i κ 2 ] A D 1 A ϕ + i D 2 2 2 A ϕ 2 + i g A | A | 2 + i G q ω A + F ,
d 2 q ( ϕ , t ) d t 2 + 2 γ d q ( ϕ , t ) d t + Ω 0 2 q ( ϕ , t ) = G ω | A ( ϕ , t ) | 2 .
a μ τ = ( κ 2 D 1 + i ζ μ ) a μ + R μ + i μ μ ( 2 δ μ μ ) a μ a μ a μ + μ μ * + f δ μ 0 ,
d 2 q ¯ d τ 2 + 2 γ ˜ d q ¯ d τ + Ω ˜ 0 2 q ¯ = η D 1 Ω 0 + 2 γ , ν a ν a ν η * e i η τ ,
R μ = i G R q ¯ μ a μ e i η τ ,
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.