This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Intra-cavity field dynamics near avoided mode crossing in concentric silicon nitride ring resonator

Maitrayee Saha [email protected] Department of Physics, Indian Institute of Technology Kharagpur, Kharagpur-721302, India    Samudra Roy Department of Physics, Indian Institute of Technology Kharagpur, Kharagpur-721302, India Centre of Theoretical Studies, Indian Institute of Technology Kharagpur, Kharagpur-721302, India    Shailendra K. Varshney Department of Electronics and Electrical Communication Engineering, Indian Institute of Technology Kharagpur, Kharagpur-721302, India
Abstract

Understanding the intra-cavity field dynamics in passive microresonator systems has already been intriguing. It becomes fascinating when the system is complex, such as a concentric dual microring resonator that exhibit avoided mode crossing (AMC). In this work, we present a systematic study of intra-cavity oscillatory field dynamics near AMC in a concentric silicon nitride microring resonator with the help of coupled Lugiato-Lefever equation (LLE). We identify two regions where the hybrid modes are strongly coupled and weakly coupled based on their eigenfrequency separation, which originate from mode coupling near AMC. We calculate the overlap integral to identify the cut-off pump wavelength region to observe the existence of the two hybrid modes. In strongly coupled hybrid mode region, we observe intra-cavity power oscillation and transfer of energy between the two hybrid modes in a periodic manner. We also evaluate the non-identical phases variation of these two modes. In weakly coupled hybrid mode region, power oscillation and energy transfer between modes reduces significantly , whereas their phase vary in almost identical fashion. We validate our numerical findings with the semi-analytical variational method, leading to an in-depth understanding of the mode coupling induced dynamics. We finally analyze the polarization properties of the field confined in the coupled system. Exploiting the Stokes parameters and Jones vector, we deduce a polarization evolving state and a polarization locked state in strongly coupled region and weakly coupled region, respectively.

pacs:
Valid PACS appear here
preprint: APS/123-QED

I Introduction

Microresonator platforms have gained considerable research interest in the last decade due to its numerous applications and scopes Brasch ; Lee . Optical frequency comb (FC) and its temporal manifestation in the form of cavity solitons (CSs) are the results of light-matter interaction in these resonator structures Hansson ; Herr . Mode-locked CS-based FC has paved the way towards precision spectroscopy Weichman , wavelength division multiplexed optical communication Fulop , optical clocks Papp , ultrafast optical ranging Trocha , optical switches Guo and so on.

Optical resonator supports multiple mode families depending on the number of total internal reflection a light ray suffers within a single round trip. These are called resonant modes (M)(M) of the system, which can be estimated from the relation 2πRnclad/λpM2πRncore/λp2\pi Rn_{\text{clad}}/\lambda_{p}\leq M\leq 2\pi Rn_{\text{core}}/\lambda_{p}, (more specifically can be defined as, M=2πRneff/λpM=2\pi Rn_{eff}/\lambda_{p} ) where, nn is the refractive index of the material and RR is the radius of the cavity Lam . A stable single mode-locked pulse dynamics can be interpreted as a double balance between Kerr nonlinearity, dispersion, the external CW pump, and the total loss in the dissipative systems Haldar ; Roy . Moreover, anomalous dispersion is desirable to obtain a bright soliton pulse and broad FC. Resonators with normal dispersion, in general, hinders the possibility of CS, where coupling between two normal dispersion modes modifies the local dispersion of one mode to the anomalous region and helps to produce a stable, bright pulse in the same resonator system Fujii ; Liu . Researchers have studied the dynamics of the resonator field in the presence of various mode coupling phenomena such as coupling between clockwise-anticlockwise modes Fujii2 ; Yoshiki , orthogonally polarized modes Saha ; Suzuki ; Hansson2 or modes which encounters avoided mode crossing (AMC) Fujii ; Liu ; Aguanno ; Lucas .
In the present work, our sole interest is to study the temporal and spectral dynamics of the resonator field in presence of mode coupling near the AMC region in a concentric microring resonator. When two modes of a resonator system have an identical resonant frequency, a new pair of hybrid modes may appear due to mode coupling between them Wiersig ; Soltani . We find out the intra-cavity field dynamics in two regions i.e. strongly coupled hybrid mode region (SCR) and weakly coupled hybrid mode region (WCR) based on the eigenfrequency separation of the two-hybrid modes Soltani . We observe oscillations of peak power as well as transfer of energy between the hybrid modes in the temporal domain and periodic variation of FC width in the spectral domain. We also witness non-identical and identical phase variations of the mode field in SCR and WCR respectively. To distinguish the pump wavelength region within which hybrid modes will exist, we calculate the overlap integral (OI) between the two hybrid mode fields. Depending on the OI values, we obtain the boundary region for strongly coupled and weakly coupled hybrid modes. Outside the boundary region for weakly coupled hybrid modes, we can safely say that hybrid modes do not exist as their resonance frequencies are too far to have sufficient amount of spatial overlap. We further verified the field dynamics theoretically with the help of Rayleigh’s dissipation function (RDF) in the framework of variational principle RoyJLT ; Saha2 ; Sahoo . We also study the polarization state of the intra-cavity field in these two separate regions with Stokes parameters and Jones vector.

II Resonator Modelling

Refer to caption
Figure 1: (a) Schematic of silicon nitride concentric microring resonator with silica cladding (b) 2D waveguide geometry of the same, W=in2600{}_{\text{in}}=2600 nm, W=out1085{}_{\text{out}}=1085 nm, g =900=900 nm, h =300=300 nm, R =50=50 μ\mum (c) Variation of effective refractive index (neff{}_{\text{eff}}) profile with wavelength (λ)(\lambda); neff{}_{\text{eff}} of individual fundamental modes of two resonators (TEinner0{}_{0}^{\text{inner}} and TEouter0{}_{0}^{\text{outer}}) crosses around 1.56 μ\mum (solid lines), mode coupling induced hybrid modes avoiding the crossing point (dotted lines)[shown in inset]. Symmetric mode makes a transition from outer to inner resonator around the AMC region as the wavelength increases and anti-symmetric mode undergoes the opposite transition.

In this section, we model and characterize concentric microresonators exhibiting avoided mode crossing Soltani ; Kim . The concentric microresonators is made of Si3N4 with SiO2 cladding, shown in Fig.1(a). The structural parameters that we consider for this waveguide are: thickness (h) of the individual ring is 300300 nm, ring radius (R) is 5050 μ\mum, widths (Win{}_{\text{in}} and Wout{}_{\text{out}}) are 26002600 nm and 10851085 nm for inner and outer rings, respectively, the gap (gg) between the two rings is 900900 nm. The design geometry is shown in Fig.1(b). It is observed that in the absence of any mode-coupling, the effective refractive index (neffn_{\text{eff}}) of the fundamental transverse electric modes (TEinner0{}_{0}^{\text{inner}} and TEouter0{}_{0}^{\text{outer}} ) for two resonators crosses each other at a particular wavelength. In Fig.1(c), we plot neffn_{\text{eff}} as a function of λ\lambda where the cross-over takes place around λ1.57\lambda\approx 1.57 μ\mum (solid lines). This AMC position can be tuned with structural parameters like gg or Win{}_{\text{in}} and Wout{}_{\text{out}}. The hybrid mode, which moves from outer resonator to inner resonator with increasing wavelength, is known as symmetric mode and the mode which undergoes a transition from inner resonator to outer one, known as asymmetric mode, respectively Kim . We demonstrate the vectorial field profile of the electric field confined in the two hybrid modes (indicated by arrow lines). It is evident that, for the symmetric mode, the electric field orientation in two coupled resonators are in the same direction whereas, for anti-symmetric modes, the orientation of the field is in opposite direction and the field also reverses its direction in either side of the AMC point. From the coupled-mode analysis for a loss-less resonator, one can readily write down the eigenfrequencies (ωas,s)(\omega^{\text{as,s}}) of these two hybrid modes as Fujii :

Refer to caption
(a)
Refer to caption
(b)
Figure 2: (a) Normalized angular resonance frequency difference ((ωasωs)/FSRouter(\omega_{as}-\omega_{s})/FSR_{\text{outer}}) of two hybrid modes vs. the pump wavelength. λ1\lambda_{1} and λ2\lambda_{2} are the lower and upper cut-off values of the SCR, shaded regions represent weakly coupled hybrid mode region (WCR). (b) Effective mode area of two hybrid modes as function of pump wavelength. (c) Variation of inverse group velocities (β1\beta_{1}) of symmetric and anti-symmetric modes with pump wavelength. (d) Group velocity dispersion (β2\beta_{2}) of two hybrid modes, two dotted lines (λ1,λ2\lambda_{1},\lambda_{2}) marked for boundary of SCR.
ωas,s=ω1+ω22±(ω1ω2)24+|κ|24.\omega^{\text{as,s}}=\frac{\omega_{1}+\omega_{2}}{2}\pm\sqrt{\frac{(\omega_{1}-\omega_{2})^{2}}{4}+\frac{|\kappa|^{2}}{4}}. (1)

Exploiting the (neffn_{\text{eff}}-λ\lambda) plot obtained from the mode analysis in COMSOL, we calculate the eigenfrequencies of the modes both in the uncoupled (ω1,ω2)(\omega_{1},\omega_{2}) and coupled conditions using the following relation ω=2πcneffλ\omega=\frac{2\pi c}{n_{\text{eff}}\lambda}. Thus, we readily obtain the mode coupling value (κ)(\kappa) from Eq. (1). The resonant modes have slightly different coupling factor i.e., κ1κ2\kappa_{1}\neq\kappa_{2} because the concentric ring waveguides are slightly different in size. We further use coupling coefficient (κ)(\kappa) : κ=κ1κ2\kappa=\sqrt{\kappa_{1}\kappa_{2}} as defined in Okamoto . In Fig.2(a), we demonstrate the resonance splitting ΔωR=(ωasωs)\Delta\omega_{R}=(\omega_{as}-\omega_{s}) of two hybrid modes which is normalized by the free spectral range (FSRouterFSR_{\text{outer}}) of the outer ring resonator. We mark two regions for hybrid modes namely, (i) strongly coupled hybrid mode region (SCR) (1.53 μ\mum λ\leq\lambda\leq 1.61 μ\mum) and (ii) weakly coupled hybrid mode region (WCR)(1.49 μ\mum λ<\leq\lambda< 1.53 μ\mum and 1.61 μ\mum <λ<\lambda\leq 1.67 μ\mum). In SCR, the two hybrid modes are correlated with the eigenmodes of each microring and can exist only when ΔωR/FSRouter1\Delta\omega_{R}/FSR_{\text{outer}}\leq 1. For WCR, 1<ΔωR/FSRouter<21<\Delta\omega_{R}/FSR_{\text{outer}}<2, indicating reduced coupling effect between the hybrid modes. In Fig.2(b), we demonstrate the variation of effective mode area (AeffA_{\text{eff}}) for both hybrid modes as a function of pump wavelength. Note that, AeffA_{\text{eff}} attains a peak value near AMC wavelength. In Fig.2(c), we plot the variation of group velocity inverse for two modes, β1(as,s)=vg(as,s)1\beta_{\text{1(as,s)}}=v_{\text{g(as,s)}}^{-1}, where vg(as,s)v_{\text{g(as,s)}} represent the group velocities of corresponding hybrid modes. It is observed that at AMC wavelength, both the modes have nearly identical β1\beta_{1} values. In Fig.2(d), we demonstrate the group velocity dispersion (GVD) profile for symmetric and anti-symmetric modes. It is interesting to note that under mode coupling condition, symmetric mode always remains in normal GVD regime. Conversely, in the SCR the anti-symmetric mode undergoes anomalous GVD regime. Vertical dotted lines mark the lower and upper cutoff values of the SCR.

Refer to caption
Figure 3: (a) Overlap integral between two hybrid modes as a function of pump wavelength. Cut-off wavelength region for strongly coupled hybrid modes and weakly coupled hybrid modes are indicated, AMC wavelength is marked with dotted vertical line around 1.57 μ\mum. Gradient shaded region indicates nearly independent eigenmodes where the mode coupling effect is negligible. Vectorial anti-symmetric hybrid mode field distribution is shown at different region.

We calculate the overlap integral (OI) between the two hybrid mode fields exploiting the following expression Damask ,

OI=Sψas(x,y).ψs(x,y)dxdyS|ψas(x,y)|2𝑑x𝑑yS|ψs(x,y)|2𝑑x𝑑yOI=\frac{\iint_{S}\psi_{as}(x,y).\psi_{s}^{*}(x,y)\hskip 2.84526ptdxdy}{\sqrt{\iint_{S}|\psi_{as}(x,y)|^{2}\hskip 2.84526ptdxdy\iint_{S}|\psi_{s}(x,y)|^{2}\hskip 2.84526ptdxdy}} (2)

The information of the spatial distribution of the hybrid modes (ψas,s(x,y)\psi_{as,s}(x,y)) are obtained from COMSOL. In Fig.(3), we plot the OI values as a function of wavelength and identify the region where the hybrid modes exist (1.49 μ\mum λ<1.67\leq\lambda<1.67 μ\mum). When OI \geq 0.55, we found strong mode coupling effect between the hybrid modes. In the limit, 0.30 \leq OI << 0.55, intracavity dynamics follows the characteristics of WCR. The spatial overlap between the two hybrid modes are negligible for OI << 0.30. In this region, the individual resonance frequencies are so far apart that the concept of hybrid modes nearly vanishes. Dotted vertical line around λ1.57\lambda\approx 1.57 μ\mum signifies the position of AMC wavelength. We have also illustrate the vectorial field distribution for anti-symmetric hybrid mode in different region.

In order to study the intra-cavity field dynamics, we obtain the fundamental mode (ωl0)(\omega_{l0}) family whose eigenfrequency is nearest to the pump frequency (ωp)(\omega_{p}). For two hybrid modes with different neffn_{\text{eff}}, the fundamental pump mode number (l0(as,s))(l_{0\text{(as,s)}}) and their eigenfrequency (ωl0)(\omega_{l0}) will be slightly different (l0(as,s)=neff(as,s)ωpRc)(l_{0\text{(as,s)}}=\frac{n_{\text{eff(as,s)}}\omega_{p}R}{c}) for a single ωp\omega_{p}. One can expand the eigenfrequencies of the fundamental mode family (l)(l) using a Taylor series expansion to obtain for first N elements Grelu :

ωl=ωl0(as,s)+k=1Nζkk!(ll0(as,s))k,\omega_{l}=\omega_{l0(\text{as,s})}+\displaystyle\sum_{k=1}^{N}\frac{\zeta_{k}}{k!}(l-l_{0(\text{as,s})})^{k}, (3)

where ωl0(as,s)\omega_{l0\text{(as,s)}} is the eigenfrequency of the individual pump mode (nearest to the external pump frequency) and ζk=dkωdlk|l=l0,(as,s)\zeta_{k}=\frac{d^{k}\omega}{dl^{k}}|_{l=l_{0,\text{(as,s)}}}. Now, ζ1=dωdl|l=l0,(as,s)=ΔωFSR\zeta_{1}=\frac{d\omega}{dl}|_{l=l_{0,\text{(as,s)}}}=\Delta{\omega_{\text{FSR}}} is the FSR of the corresponding resonator supporting the hybrid mode at pump frequency. The resonator being dispersive, the eigenmodes are not exactly equidistant, ζ2=d2ωdl2|l=l0,(as,s)\zeta_{2}=\frac{d^{2}\omega}{dl^{2}}|_{l=l_{0,\text{(as,s)}}} is the second order dispersion coefficient for the hybrid modes. ζ2\zeta_{2} can be related to the well-known notation for GVD (β2\beta_{2}) as, ζ2=cneffζ12β2\zeta_{2}=-\frac{c}{n_{\text{eff}}}\zeta_{1}^{2}\beta_{2} Grelu . We study field dynamics at different pump frequencies near the AMC region. At each time, resonant mode number changes with external pump frequency and correspondingly ωl0(as,s)\omega_{l0(\text{as,s})}, ΔωFSR\Delta{\omega_{\text{FSR}}}, dispersion coefficients are calculated from the Taylor expansion coefficients.

III Coupled Lugiato-Lefever Equation Analysis

The field dynamics inside a dissipative system like Kerr microresonator can be modelled by the well-known Lugiato-Lefever equation (LLE) Lugiato , which incorporates the Kerr nonlinearity, dispersion, external pump source and the total loss of the system. Apart from all these parameters, this system being dissipative, the field dynamics is highly sensitive to the external parameters like frequency detuning, i.e., the difference between pump frequency and the nearest cavity mode resonant frequency (ωpωl0(as,s)\omega_{p}-\omega_{l0(\text{as,s})}). Here, we use coupled LLE to find the dynamics of hybrid modes experiencing the AMC. The coupled LLE’s are as follows:

uast=(12ΔΓas+iσas+Δζ1θ+iζ2as2!2θ2+ig0as(|uas|2+2|us|2))uas+iκ2us+12ΔΓasFas,\frac{\partial u_{as}}{\partial t}=\Bigl{(}-\frac{1}{2}\Delta\Gamma_{as}+i\sigma_{as}+{\color[rgb]{0,0,0}\Delta\zeta_{1}\frac{\partial}{\partial\theta}}+i\frac{\zeta_{2_{as}}}{2!}\frac{\partial^{2}}{\partial\theta^{2}}\\ +ig_{0as}(|u_{as}|^{2}+{\color[rgb]{0,0,0}2|u_{s}|^{2})}\Bigr{)}u_{as}+i\frac{\kappa}{2}u_{s}+\frac{1}{2}\Delta\Gamma_{as}F_{as}, (4)
ust=(12ΔΓs+iσsΔζ1θ+iζ2s2!2θ2+ig0s(|us|2+2|uas|2))us+iκ2uas+12ΔΓsFs.\frac{\partial u_{s}}{\partial t}=\Bigl{(}-\frac{1}{2}\Delta\Gamma_{s}+i\sigma_{s}-{\color[rgb]{0,0,0}\Delta\zeta_{1}\frac{\partial}{\partial\theta}}+i\frac{\zeta_{2_{s}}}{2!}\frac{\partial^{2}}{\partial\theta^{2}}\\ +ig_{0s}(|u_{s}|^{2}+{\color[rgb]{0,0,0}2|u_{as}|^{2})}\Bigr{)}u_{s}+i\frac{\kappa}{2}u_{as}+\frac{1}{2}\Delta\Gamma_{s}F_{s}. (5)

In the construction of coupled LLE, Eq.(4) and Eq.(5) uas{u_{as}} and us{u_{s}} represent the field amplitudes of anti-symmetric and symmetric modes, respectively. In this scenario, either one of the hybrid modes (anti-symmetric or symmetric) is pumped at any particular pump frequency depending on, which mode confines in the outer microring resonator. At the lower pump frequency, anti-symmetric mode is excited (Fas0,Fs=0)(F_{as}\neq 0,F_{s}=0) whereas for higher pump frequency, symmetric mode is excited (Fas=0,Fs0)(F_{as}=0,F_{s}\neq 0). The coupling coefficient (κ\kappa) arises due to the degeneracy in eigenfrequency of two different resonator eigenmodes. Here, κ\kappa is equivalent to the separation between the two hybrid modes, which varies with pumping wavelength near the AMC region suggesting the linear coupling effect between these two hybrid modes. Due to the sufficient amount of overlapping field between these modes near AMC, we also include nonlinear cross-phase modulation (XPM) effect (sixth term on the right hand side of coupled equations (Eqs.4-5)). Though, for the choice of the external parameters in this case, we have seen the effect of XPM is negligible compare to the effect of linear coupling between the two modes (see, Appendix 11). We further neglect any coherent coupling terms. It is possible when light is confined within the resonator for a long time covering multiple round trips, such that the traverse path length is much greater than the corresponding beat length Agrawal . Here, tt is the slow time, θ\theta is the azimuthal coordinate of the resonator, ΔΓas,s\Delta\Gamma_{\text{as,s}} is the resonance linewidth of a mode which can be determined from the relation ΔΓas,s=ωl0(as,s)Q\Delta\Gamma_{\text{as,s}}=\frac{\omega_{l0\text{(as,s)}}}{Q}. The Q-factor of a cavity can be calculated as Q=ωpTrtlrtQ=\frac{\omega_{p}T_{\text{rt}}}{l_{\text{rt}}} where Trt=2πζ11.98T_{\text{rt}}=\frac{2\pi}{\zeta_{1}}\simeq 1.98 ps is the round trip time and lrtl_{\text{rt}} is the round trip loss. Round trip number can be estimated as, ‘total time confined/ round trip time’=t/Trtt/T_{rt}. The nonlinear coefficient g0g_{0} is defined as, g0(as,s)=ωl0(as,s)2n2ζ1(as,s)2πn2Aeff(as,s)g_{0\text{(as,s)}}=\frac{\hbar\omega_{l0\text{(as,s)}}^{2}n_{2}\zeta_{1\text{(as,s)}}}{2\pi n^{2}A_{\text{\text{eff(as,s)}}}}, \hbar, n2n_{2}, nn, Aeff(as,s)A_{eff(as,s)} are the reduced Planck constant, Kerr coefficient, the linear refractive index of the medium and effective area of the mode, respectively. κ\kappa is the mode coupling constant, Fas,sF_{\text{as,s}} is the external pump field defined as (1/2)ΔΓas,sFas,s=ΔΓas,sPinωl0,(as,s)(1/2)\Delta\Gamma_{\text{as,s}}F_{\text{as,s}}=\sqrt{\Delta\Gamma_{\text{as,s}}}\sqrt{\frac{P_{\text{in}}}{\hbar\omega_{l0,\text{(as,s)}}}}, where PinP_{\text{in}} is the input power. We observe linewidth of the hybrid modes are of the same order, i.e. ΔΓasΔΓs1\frac{\Delta\Gamma_{as}}{\Delta\Gamma_{s}}\approx 1. So, from now onward we designate ΔΓasΔΓs=ΔΓ\Delta\Gamma_{as}\approx\Delta\Gamma_{s}=\Delta\Gamma 50 GHz\approx 50\text{ GHz}. σas,s=ωpωl0(as,s)\sigma_{\text{as,s}}=\omega_{p}-\omega_{l0\text{(as,s)}} is the detuning between the pump and resonance frequencies of the resonator. Δζ1=ζ1,asζ1,s\Delta\zeta_{1}=\zeta_{1,as}-\zeta_{1,s} is the difference between cavity FSR for the two hybrid modes. The spectral variation of FSR for the hybrid modes follow the similar pattern of β1\beta_{1} (Fig.2(c)), mentioned in Aguanno . For numerical simulations, it is useful to normalize the coupled LLE which takes the form,

ψasτ=(1iαas+D1θ+iD2as2θ2+i(|ψas|2+2|ψs|2))ψas+iχψs+Sas,\frac{\partial\psi_{as}}{\partial\tau}=\Bigl{(}-1{\color[rgb]{0,0,0}-i\alpha_{as}+D_{1}\frac{\partial}{\partial\theta}}+iD_{2as}\frac{\partial^{2}}{\partial\theta^{2}}\\ +i(|\psi_{as}|^{2}+{\color[rgb]{0,0,0}2|\psi_{s}|^{2})}\Bigr{)}\psi_{as}+i\chi\psi_{s}+S_{as}, (6)
ψsτ=(1iαsD1θ+iD2s2θ2+i(|ψs|2+2|ψas|2))ψs+iχψas+Ss,\frac{\partial\psi_{s}}{\partial\tau}=\Bigl{(}-1{\color[rgb]{0,0,0}-i\alpha_{s}-D_{1}\frac{\partial}{\partial\theta}}+iD_{2s}\frac{\partial^{2}}{\partial\theta^{2}}\\ +i(|\psi_{s}|^{2}+{\color[rgb]{0,0,0}2|\psi_{as}|^{2})}\Bigr{)}\psi_{s}+i\chi\psi_{as}+S_{s}, (7)

where the parameters are rescaled as : ψas=2g0asΔΓuas\psi_{as}=\sqrt{\frac{2g_{0as}}{\Delta\Gamma}}u_{as}, ψs=2g0sΔΓus\psi_{s}=\sqrt{\frac{2g_{0s}}{\Delta\Gamma}}u_{s}, αas,s=2σas,sΔΓ\alpha_{\text{as,s}}=-\frac{2\sigma_{\text{as,s}}}{\Delta\Gamma}, D2(as,s)=ζ2(as,s)ΔΓD_{2\text{(as,s)}}=\frac{\zeta_{2\text{(as,s)}}}{\Delta\Gamma}, D1=2Δζ1ΔΓD_{1}=\frac{2\Delta\zeta_{1}}{\Delta\Gamma}, χ=κΔΓ\chi=\frac{\kappa}{\Delta\Gamma}, Sas,s=4g0(as,s)PinΔΓ2ωpS_{\text{as,s}}=\sqrt{\frac{4g_{0\text{(as,s)}}P_{\text{in}}}{\Delta\Gamma^{2}\hbar\omega_{p}}} and τ=ΔΓ2t\tau=\frac{\Delta\Gamma}{2}t. For simplicity, in simulations, we have excluded all higher-order dispersion and nonlinearity terms. We choose our initial wave-function of the form,

ψas,s=ηas,ssech(ηas,sθ),\psi_{as,s}=\eta_{as,s}\operatorname{sech}\Bigl{(}\eta_{as,s}\theta\Bigr{)}, (8)

where, ηas,s=2|αas,s|\eta_{as,s}=\sqrt{2|\alpha_{as,s}|} and αas,s\alpha_{as,s} is the normalized detuning parameter for the two hybrid modes. For a particular pump wavelength, the difference between the eigenfrequencies of the hybrid modes are very small (ωl0(as)ωl0(s))(\omega_{l0\text{(as)}}\approx\omega_{l0\text{(s)}}), such that the detuning value for both modes are nearly same. Here, the normalized detuning value αas,s\alpha_{as,s} is considered as unity for both cases. Further, we analyze two regions of strongly and weakly coupled hybrid modes on the basis of their eigenfrequency separation. We observe damped oscillatory motion of intra-cavity field due to mode coupling for a particular choice of detuning frequency. Phase and the polarization state of the cavity field evolve significantly in a different way in these two regions.

III.1 Strongly Coupled Region

Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a and b) Temporal evolution of field and (c and d) evolution of frequency comb profile of anti-symmetric and symmetric hybrid modes, respectively, at the AMC pumped near 1.567 μ\mum.
Refer to caption
Figure 5: Periodic transfer of energy between two hybrid modes with propagation. In both cases the initial energy is 222\sqrt{2}.

Strongly coupled hybrid modes exist within a frequency window where the difference between their eigenfrequencies i.e., ΔωR=(ωasωs)\Delta\omega_{R}=(\omega_{as}-\omega_{s}) is less than the FSR of the outer ring resonator, satisfying the condition ΔωRFSRouter1\frac{\Delta\omega_{R}}{FSR_{\text{outer}}}\leq 1, shown in Fig.2(a). The main characteristics of SCR are; i) both hybrid modes show opposite dispersion i.e., the anti-symmetric mode falls in the anomalous dispersion while the symmetric mode experience normal dispersion. Further, the GVD attains its peak value (+ve{}^{\prime}+ve^{\prime} for symmetric, ve{}^{\prime}-ve^{\prime} for anti-symmetric) at AMC (see Fig.2(d)), ii) GVM (δ=|1/β1,as1/β1,s|)(\delta=|1/\beta_{1,as}-1/\beta_{1,s}|) of two hybrid modes is minimum at AMC. In Fig.(4), we demonstrate the overall field dynamics at the AMC by numerically solving the coupled LLE. The parameters obtained from the structure at this pump wavelength are: ζ1,as=3.1771\zeta_{1,as}=3.1771 THz, ζ2,as=1.714\zeta_{2,as}=1.714 GHz, ζ1,s=3.1720\zeta_{1,s}=3.1720 THz, ζ2,s=3.420\zeta_{2,s}=-3.420 GHz, κ=1.024\kappa=1.024 THz. We use cw pump power Pin=0.5P_{\text{in}}=0.5 W to the system. The resonator eigenfrequency near the pump mode (ωl0(as,s))(\omega_{l0\text{(as,s)}}) is obtained from Eq.(3). At the AMC wavelength, we obtain ωl0(as)=1.20315×1015\omega_{l0(as)}=1.20315\times 10^{15}Hz, ωl0(s)=1.20247×1015\omega_{l0(s)}=1.20247\times 10^{15}Hz. In Fig.4((a)-(b)), we depict the normalized peak power evolution for temporal field for both symmetric and anti-symmetric modes at the avoided crossing point. The spectral counterpart showing the frequency comb width also periodically varies with propagation, shown in Fig.4((c)-(d)). We observe the peak power confined in each mode gradually decays with round trip time but with an oscillatory fashion. This periodic peak power evolution is not due to the breathing pulsation rather due to periodic energy exchange between modes. In order to confirm the periodic energy exchange we numerically calculate the total energy confined in each hybrid mode as, E=|ψas/s|2𝑑θE=\int_{-\infty}^{\infty}|\psi_{as/s}|^{2}\hskip 2.84526ptd\theta. In Fig.(5), we plot the evolution of the total energy of the symmetric and anti-symmetric modes. Initially, the energy confined in both hybrid modes are equal, with propagation there is a periodic exchange of energy between them. In both the cases (symmetric and anti-symmetric) the overall energy decays with round trip exhibiting the transient nature of the mode in AMC region.

Refer to caption
Figure 6: (a) Power oscillations between symmetric and anti-symmetric modes at two edge pump wavelength (λ=1.53\lambda=1.53 μ\mum, λ=1.6\lambda=1.6 μ\mum) of strongly coupled region (SCR). (b) Variation of oscillation period (τos)(\tau_{os}) with pump wavelength within SCR for anti-symmetric mode (blue hollow circle) and symmetric mode (pink filled circle), solid curve shows the absolute GVM (δ\delta) variation between the hybrid modes with the pump wavelength.

In Fig.6(a), we numerically demonstrate the evolution of the peak power (PP) of two modes with slow time (τ\tau) at two extreme wavelengths of SCR. We observe that these two modes’ power remains correlated even at the edge wavelengths in either side of AMC, but oscillate with different oscillation periods. The oscillation period varies with the pump wavelength λp\lambda_{p}. In Fig.6(b), we map the oscillation period τos\tau_{os} for anti-symmetric (blue hollow circles) and symmetric (pink filled circles) modes field with operating wavelength and notice that τos\tau_{os} attends its minima near the AMC wavelength (λAMC\lambda_{AMC}). We further find that the variation of τos\tau_{os} closely follows the nature of absolute GVM (δ=|1β1,as1β1,s|\delta=|\frac{1}{\beta_{1},as}-\frac{1}{\beta_{1},s}|) between these two hybrid modes. Like τos\tau_{os}, δ\delta attends its minima at λAMC\lambda_{AMC} and increases in either sides of λAMC\lambda_{AMC}. This correspondence leads us to the conclusion that GVM plays a critical role during the field oscillations.

III.2 Weakly Coupled Region

Refer to caption
(a)
Refer to caption
(b)
Figure 7: (a and b) Temporal and (c and d) spectral profile of anti-symmetric and symmetric hybrid modes, respectively, pumped at 1.614 μ\mum in weakly coupled region.

In this subsection, we discuss the basic characteristics of the hybrid modes when their coupling strength gradually reduces. This weakly coupled region lies for the pump wavelength 1.49 μ\mum λ<\leq\lambda< 1.53 μ\mum and 1.61 μ\mum <λ<\lambda\leq 1.67 μ\mum as depicted by the shaded region in Fig.2(a). We characterize the WCR which satisfies the two following conditions: i) where, ΔωR\Delta\omega_{R} satisfies the condition 1<ΔωRFSRouter<21<\frac{\Delta\omega_{R}}{FSR_{\text{outer}}}<2 and ii) overlap integral (OI) between the two hybrid modes lies within the range, 0.3OI<0.550.3\leq OI<0.55. In Fig.(7), we demonstrate the temporal and spectral characteristics of hybrid modes when pumped in WCR (λp=1.614(\lambda_{p}=1.614 μ\mum). The parameters used in the simulation are as follows: ζ1,as=3.1433\zeta_{1,as}=3.1433 THz, ζ2,as=1.7154\zeta_{2,as}=1.7154 GHz, ζ1,s=3.2456\zeta_{1,s}=3.2456 THz, ζ2,s=3.2577\zeta_{2,s}=-3.2577 GHz, κ=1.174\kappa=1.174 THz, cw pump power used here same as before (0.5W). Now, depending on the pump wavelength, whichever hybrid mode stays in the outer resonator gets directly pumped (Sas/s0S_{\text{as/s}}\neq 0) and the other mode is evanescently coupled to the mode present in the inner resonator (Sas/s=0S_{\text{as/s}}=0). Here, We also obtain the eigenfrequencies (ωl0(as,s))(\omega_{l0\text{(as,s)}}) using Eq.(3). At λp=1.614\lambda_{p}=1.614 μ\mum, we obtain ωl0(as)=1.167441×1015\omega_{l0(as)}=1.167441\times 10^{15}Hz, ωl0(s)=1.167487×1015\omega_{l0(s)}=1.167487\times 10^{15}Hz. In Fig.7(a-b) and Fig.7(c-d), we represent the temporal and spectral characteristics of anti-symmetric and symmetric modes respectively for the aforementioned parameters. Unlike the situation at AMC (see Fig.(4)), temporal field evolution is i) asymmetrical as the GVM increases, the field is either gradually ahead (anti-symmetric mode) or delayed (symmetric mode) from its initial position, ii) periodic exchange of energy between the hybrid modes gradually reduces and decays nearly in an identical way with propagation. We observe similar asymmetries in the spectral domain as well due to the increased GVM. Either left (anti-symmetric mode) or right (symmetric mode) wing width of the frequency comb oscillates with propagation. In the other side of WCR, i.e. for 1.49 μ\mum λ<\leq\lambda< 1.53 μ\mum, the only difference in the field characteristics from the one that are reported in Fig.(7) is the nature of asymmetries in temporal and spectral profiles. The position of the asymmetries will be reversed due to the opposite GVM faced by two-hybrid modes in this side of AMC point.

IV Variational analysis

In this section, we introduce the Lagrangian perturbative approach based on variational principle Anderson ; RoyJLT to analyze the field dynamics semi-analytically. The treatment requires a suitable ansatz whose parameters may evolve with slow time. Note, a major approximation of this treatment is the intactness of the ansatz function. We proceed with standard Kerr soliton ansatz having the mathematical form:

ψas=ηassech(ηasθ)ei(ϕas+D1θ)\psi_{as}=\eta_{as}\text{sech}(\eta_{as}\theta)e^{i(\phi_{as}+D_{1}\theta)} (9)
ψs=ηssech(ηsθ)ei(ϕsD1θ),\psi_{s}=\eta_{s}\text{sech}(\eta_{s}\theta)e^{i(\phi_{s}-D_{1}\theta)}, (10)

where the amplitudes (ηas,ηs)(\eta_{as},\eta_{s}) and the phases (ϕas,ϕs)(\phi_{as},\phi_{s}) of two CSs are allowed to vary over slow time, τ\tau. D1D_{1} is proportional to the group-velocity mismatch between the two hybrid modes defined as, D1δ/LrD_{1}\approx\delta/L_{r}, where LrL_{r} is the resonator length and δ\delta is the group-velocity mismatch between the two hybrid modes. As a standard procedure we define the Lagrangian in such a way that the following Euler-Lagrangian (EL) equation retrieve the governing Eq.(6) and Eq.(7). Note, for mathematical simplicity, we ignore the negligible effect of nonlinear coupling terms in Lagrangian formalism.

ddτ(Lψτ)+ddθ(Lψθ)Lψ+(Rψτ)=0,\frac{d}{d\tau}\left(\frac{\partial L}{\partial\psi_{\tau}^{*}}\right)+\frac{d}{d\theta}\left(\frac{\partial L}{\partial\psi_{\theta}^{*}}\right)-\frac{\partial L}{\partial\psi^{*}}+\left(\frac{\partial R}{\partial\psi_{\tau}^{*}}\right)\\ =0, (11)

where L and R consist of:

L=Las+Ls,L=L_{as}+L_{s}, (12)
R=Ras+Rs.R=R_{as}+R_{s}. (13)

Here (LasL_{as}, RasR_{as}) and (LsL_{s}, RsR_{s}) represent the Lagrangian function and RDF of the anti-symmetric and symmetric modes, respectively. Next, we reduce the Lagrangian and RDF as, Lg=L𝑑θL_{g}=\int_{-\infty}^{\infty}L\hskip 2.84526ptd\theta and Rg=R𝑑θR_{g}=\int_{-\infty}^{\infty}R\hskip 2.84526ptd\theta (details are given in Appendix B). Now we use the reduced form of LgL_{g} and RgR_{g} in EL equation:

ddτ(Lgpj˙)Lgpj+(Rgpj˙)=0,\frac{d}{d\tau}\left(\frac{\partial L_{g}}{\partial\dot{p_{j}}}\right)-\frac{\partial L_{g}}{\partial p_{j}}+\left(\frac{\partial R_{g}}{\partial\dot{p_{j}}}\right)=0, (14)

where pj=ηas,ηs,ϕas,ϕsp_{j}=\eta_{as},\eta_{s},\phi_{as},\phi_{s} and pj˙=ηasτ,ηsτ,ϕasτ,ϕsτ\dot{p_{j}}=\frac{\partial\eta_{as}}{\partial\tau},\frac{\partial\eta_{s}}{\partial\tau},\frac{\partial\phi_{as}}{\partial\tau},\frac{\partial\phi_{s}}{\partial\tau}, we obtain the coupled ODEs for two hybrid mode field parameters. The coupled differential equations governing the amplitude and phase of these fields are as follows:

Refer to caption
Figure 8: (a) Peak power and (b) phase variation for symmetric and anti-symmetric mode shown both numerically (solid lines) and with variational analysis (dotted lines) pumped at λ=1.567\lambda=1.567 μ\mum in SCR.
ηasτ=2ηasχηsΔassinϕcsch(Δas)¯+Sasπcosϕassech(Δas2),\frac{\partial\eta_{as}}{\partial\tau}={\color[rgb]{0,0,0}-2\eta_{as}}\underline{-\chi\eta_{s}\Delta_{as}\sin\phi\operatorname{csch}\left(\Delta_{as}\right)}\\ +S_{as}\pi\cos\phi_{as}\operatorname{sech}\left(\frac{\Delta_{as}}{2}\right), (15)
ηsτ=2ηs+χηasΔssinϕcsch(Δs)¯+Ssπcosϕssech(Δs2),\frac{\partial\eta_{s}}{\partial\tau}={\color[rgb]{0,0,0}-2\eta_{s}}\underline{+\chi\eta_{as}\Delta_{s}\sin\phi\operatorname{csch}\left(\Delta_{s}\right)}\\ +S_{s}\pi\cos\phi_{s}\operatorname{sech}\left(\frac{\Delta_{s}}{2}\right), (16)
ϕasτ=Δas2Sas2D1sinϕassech(Δas2)tanh(Δas2)αas+D12(1D2as)+ηas2(1D2as)χΔasηsηascosϕcsch(Δas)(1Δascoth(Δas)),\frac{\partial\phi_{as}}{\partial\tau}={\color[rgb]{0,0,0}-\frac{\Delta_{as}^{2}S_{as}}{2D_{1}}\sin\phi_{as}\operatorname{sech}\left(\frac{\Delta_{as}}{2}\right)\tanh\left(\frac{\Delta_{as}}{2}\right)}\\ -\alpha_{as}+{\color[rgb]{0,0,0}D_{1}^{2}\left(1-D_{2as}\right)}+\eta_{as}^{2}\left(1-D_{2as}\right)\\ -\chi\frac{\Delta_{as}\eta_{s}}{\eta_{as}}\cos\phi\operatorname{csch}\left(\Delta_{as}\right)\left(1-\Delta_{as}\coth\left(\Delta_{as}\right)\right), (17)
ϕsτ=Δs2Ss2D1sinϕssech(Δs2)tanh(Δs2)αs+D12(1D2s)+ηs2(1D2s)χΔsηasηscosϕcsch(Δs)(1Δscoth(Δs)),\frac{\partial\phi_{s}}{\partial\tau}={\color[rgb]{0,0,0}-\frac{\Delta_{s}^{2}S_{s}}{2D_{1}}\sin\phi_{s}\operatorname{sech}\left(\frac{\Delta_{s}}{2}\right)\tanh\left(\frac{\Delta_{s}}{2}\right)}\\ -\alpha_{s}+{\color[rgb]{0,0,0}D_{1}^{2}\left(1-D_{2s}\right)}+\eta_{s}^{2}\left(1-D_{2s}\right)\\ -\chi\frac{\Delta_{s}\eta_{as}}{\eta_{s}}\cos\phi\operatorname{csch}\left(\Delta_{s}\right)\left(1-\Delta_{s}\coth\left(\Delta_{s}\right)\right), (18)
Refer to caption
Figure 9: (a) Peak power and (b) phase variation for symmetric and anti-symmetric mode shown both numerically (solid lines) and with variational analysis (dotted lines) pumped at λ=1.614\lambda=1.614 μ\mum in WCR.
Table 1: Polarized flux components evolution with normalized slow time in weakly and strongly coupled region.
Pump wavelength (λp)\lambda_{p}) Measured Quantity Normalized slow time (τ\tau)
τ=0\tau=0 τ=0.25\tau=0.25 τ=0.50\tau=0.50 τ=0.75\tau=0.75 τ=1.0\tau=1.0
1.567 μ\mum (SCR) Degree of Linear polarization (DoLP) (S12+S22S0)\Bigl{(}\frac{\sqrt{S_{1}^{2}+S_{2}^{2}}}{S_{0}}\Bigr{)} 1 0.80 0.54 0.92 0.87
Degree of Circular polarization (DoCP) (|S3|S0)\Bigl{(}\frac{|S_{3}|}{S_{0}}\Bigr{)} 0 0.6 0.84 0.39 0.48
1.614 μ\mum (WCR) Degree of Linear polarization (DoLP) (S12+S22S0)\Bigl{(}\frac{\sqrt{S_{1}^{2}+S_{2}^{2}}}{S_{0}}\Bigr{)} 1 0.999 0.998 0.998 0.999
Degree of Circular polarization (DoCP) (|S3|S0)\Bigl{(}\frac{|S_{3}|}{S_{0}}\Bigr{)} 0 0.004 0.005 0.007 0.005

where, Δas=πD1ηas\Delta_{as}=\frac{\pi D_{1}}{\eta_{as}}, Δs=πD1ηs\Delta_{s}=\frac{\pi D_{1}}{\eta_{s}}, ϕ=ϕsϕas\phi=\phi_{s}-\phi_{as}. These set of four coupled ordinary differential equations (Eq.15-Eq.18) capture the evolution dynamics of the amplitude and phase of the anti-symmetric (ηas,ϕas)(\eta_{as},\phi_{as}) and symmetric (ηs,ϕs)(\eta_{s},\phi_{s}) modes respectively. In Fig.(8), we have shown the field characteristics when pumped in SCR at AMC wavelength (λ=1.567μ\lambda=1.567\mum). Fig.8(a) shows the normalized peak power evolution for both hybrid modes. Simulation results (solid lines) are obtained by solving coupled LLE numerically. The results show how the power of the modes follows a damped oscillatory evolution. The symmetric and anti-symmetric modes accompany an inverse phase relationship dictating switching dynamics. We compare this simulation results with the theoretical prediction made through variational analysis and find an excellent agreement. The underlined terms containing the coupling factor (χ\chi) in Eq.(15) and Eq.(16) responsible for the oscillatory pattern of the power evolution. The opposite sign of χ\chi in two equations leads to the reverse patterns of oscillation between the two modes. The total amplitude η(=ηas+ηs)\eta(=\eta_{as}+\eta_{s}) approximately follow an exponential decay as, ηη0eτ\eta\approx\eta_{0}e^{-\tau}, when η0=ηas(0)+ηs(0)\eta_{0}=\eta_{as}(0)+\eta_{s}(0). Hence, for the initial wave-function given in Eq. (8), we can have the total energy evolution ET(τ)42e2τE_{T}(\tau)\approx 4\sqrt{2}e^{-2\tau}, where ET=Eas+EsE_{T}=E_{as}+E_{s}. In Fig.8(b), we demonstrate the phase-evolution for symmetric and anti-symmetric modes. Variational results (dotted lines) support the numerical data (solid lines) to a good extent. Earlier, this non-identical phase variation near AMC has also been mentioned in Haus . Note, we wrapped the phase value within the range [π,π][-\pi,\pi]. In Fig.9(a,b), we plot the evolution of normalized peak power and phase with slow time, both numerically (solid lines) and with variational results (dotted lines), when the pump wavelength is in WCR ( λp=1.614\lambda_{p}=1.614 μ\mum). Numerical simulations reveal power oscillation to some extent within these hybrid modes; however, peak power follows exponential decay. In contrast to Fig.8(a), here, the symmetric mode decays faster than the anti-symmetric mode initially, and after which, both of them decay nearly in a similar fashion. Variational results could not capture those power oscillations; instead, it results in exponential fall of amplitudes similar to the simulation results’ envelope.

Refer to caption
Figure 10: Polarization ellipse evolution with slow time (τ\tau). (a) when the pump is in SCR at λ=1.567\lambda=1.567 μ\mum, helicity reverses at τ=0.50\tau=0.50, (b) when the pump is in WCR at λ=1.614\lambda=1.614 μ\mum, helicity maintains throughout the evolution.

Another striking contrast between the two regions is the evolution of pulse phase (ϕas,s)(\phi_{\text{as,s}}), shown in Fig.8(b) and Fig.9(b). In SCR (see Fig.8(b)), we found that the phase of the two-hybrid modes behaves non-identically, whereas, in WCR, phase evolution are nearly identical. We observe that numerical (solid lines) and variational results (dotted lines) differ quantitatively, but qualitatively both reflect the identical nature of hybrid modes’ phase variation.

V Polarization state of intra-cavity field

So far, we have seen the evolution of non-identical and identical phase variation for the two hybrid modes in SCR and WCR, respectively. Initially, we launch a linearly polarized light, which is coupled to these two hybrid modes in presence of mode coupling. In SCR, the relative phase difference between two field also changes with slow time. Thus, it is possible that the net field within the resonator created in the presence of the interference between the two hybrid mode field may experience a change from its initial polarization state. To find the polarization properties, we evaluate Stokes parameters Averlant and corresponding Jones vector Chipman . We numerically find out the S-parameters as: S0=|ψas|2+|ψs|2S_{0}=|\psi_{as}|^{2}+|\psi_{s}|^{2}, S1=|ψas|2|ψs|2S_{1}=|\psi_{as}|^{2}-|\psi_{s}|^{2}, S2=ψasψs+ψasψsS_{2}=\psi_{as}^{*}\psi_{s}+\psi_{as}\psi_{s}^{*} and S3=i(ψasψsψasψs)S_{3}=i(\psi_{as}^{*}\psi_{s}-\psi_{as}\psi_{s}^{*}). Physically, these parameters signify the amount of flux radiated in space along different orientations. S0S_{0} signifies the total flux, S1S_{1} and S2S_{2} together define the total linear flux (S12+S22\sqrt{S_{1}^{2}+S_{2}^{2}}) and S3S_{3} defines how much amount of flux is circularly polarized. We further find out that the output flux is fully polarized in both SCR and WCR by satisfying the condition S0=S12+S22+S32S_{0}=\sqrt{S_{1}^{2}+S_{2}^{2}+S_{3}^{2}}. When a light beam is fully polarized, Jones vector helps to visualize the geometrical orientation of the electric field direction. Jones vector has four degrees of freedom, two indicating the amplitudes (Ax,Ay)(A_{x},A_{y}) and two correspond to their phase (Φx,Φy)(\Phi_{x},\Phi_{y}). It can be represented in terms of S-parameters Saha :

J=(S0+S12S0S12eitan1(S3S2))\begin{matrix}J=\end{matrix}\begin{pmatrix}\sqrt{\frac{S_{0}+S_{1}}{2}}\\ \sqrt{\frac{S_{0}-S_{1}}{2}}e^{i\tan^{-1}\Bigl{(}\frac{S_{3}}{S_{2}}\Bigr{)}}\end{pmatrix} (19)

where Ax=S0+S12A_{x}=\sqrt{\frac{S_{0}+S_{1}}{2}}, Ay=S0S12A_{y}=\sqrt{\frac{S_{0}-S_{1}}{2}} and (ΦxΦy)=tan1(S3S2)(\Phi_{x}-\Phi_{y})=\tan^{-1}\Bigl{(}\frac{S_{3}}{S_{2}}\Bigr{)}. Polarized flux components for two pump wavelengths, one at SCR (1.567 μ\mum), and another at WCR (1.614 μ\mum) are listed in Table IV, which we have calculated from S-parameters, obtained through the coupled LLE at θ=0\theta=0. The degree of linear polarization (DoLP) indicates how the electric field confines in one plane. Similarly, the degree of circular polarization (DoCP) characterizes how much fraction of the polarized flux is circularly polarized. We launch a linearly polarised light (DoLP = 1, DoCP = 0). Previously, in Fig.8(a) and Fig.9(a), we observe that the peak power gradually decays with slow time. Thus, the magnitude of (S0S_{0}), as well as other S-parameters, also decrease with slow time. Though the helicity of the circularly polarized light has not taken care during the evaluation of DoCP, this fact has been introduced while drawing the polarization ellipse with Jones vectors’ help. Data of individual S-parameter evolution and corresponding Jones vector have been mentioned in Appendix C. From Table IV, we observe in SCR, the polarization state evolves with slow time, and the majority of the flux evolves periodically from linear to circular and again back to the linearly dominating flux component. In contrast to that, in WCR, the initial linear polarization state remains nearly fixed, the circular polarisation component remains negligible compare to its linear component throughout the evolution. Thus in WCR, the net field remains in a polarization locked state. In Fig.(10), we have shown the evolution of polarization ellipse with the normalized slow time drawn out of the Jones vector. While pumping at 1.567 μ\mum in SCR, the helicity of circularly polarized flux transforms from right to left orientation near τ=0.50\tau=0.50, then goes back to right circular orientation (see Fig.10(a)), along with the change in dominating flux components (linear \rightarrow circular \rightarrow linear). In Fig.10(b), we observe in WCR (λ=1.614\lambda=1.614 μ\mum), the dominating flux components always remain linear, and the helicity is also maintained throughout.
Recent studies have shown stable FC generation due to nonlinear Aguanno mode coupling, as well as linear mode coupling induced multi-FSR FC generation Fujii near AMC in an optical resonator. Choosing the external parameter like detuning frequency can lead to stable multi-FSR FC or unstable oscillatory single FSR FC generation. Though, various studies Fujii ; Liu ; Soltani ; Kim discuss the stable multi-FSR region, the dynamics in the oscillatory region has gained barely any attention. We find an experimental observation has been reported with power oscillations between two counter-propagating modes in a whispering gallery mode resonator Yoshiki , which is close to our observed damped oscillation phenomena.

VI conclusion

To conclude our work, we have reported the intra-cavity field dynamics in a concentric dual microring resonator system using coupled Lugiato-Lefever equation. We analyze mode coupling induced temporal and spectral field dynamics in the proximity of avoided mode crossing regions, namely in strongly coupled and weakly coupled hybrid mode regions, based on the their resonance frequency splitting. We further support the numerical findings with a semi-analytical variational treatment that justifies the observation of peak power oscillations or the energy transfer between the hybrid modes solely occurs due to linear mode coupling. The oscillation period varies with pump frequencies within the SCR, which seems to follow the nature of group velocity mismatch between the two hybrid modes. We observe that mode coupling induced phase-shift leads to a striking difference in polarization properties of the total intracavity field in these two regions. The non-identical phase evolution of two hybrid mode fields in SCR leads to a polarization evolving state for the total intra-cavity field. While pumping in WCR, the intra-cavity field nearly maintains its initial polarization state throughout the evolution. We also calculated the overlap integral factor between two hybrid modes to identify the region of their existence. We believe that our findings will enrich the understanding of mode coupling induced intra-cavity field dynamics in coupled micro-resonator system.

Acknowledgment

The author M.S would like to thank Ministry of Human Resource Development, Government of India and IIT Kharagpur for funding to carry out her research work. Author SKV acknowledge the support received from project vide sanction no. DST/NM/NNETRA/2018(6)-IITKGP.

Appendix A Comparison of linear and nonlinear mode coupling effect on pulse dynamics

We have included linear and nonlinear cross-phase modulation (XPM) effect in the coupled LLE (Eq.4-Eq.5) to model hybrid mode evolution in the resonator. Here, we have used the XPM coefficient value to be ‘2’ Aguanno . In this section, we have shown the individual linear and nonlinear coupling effect in Fig.(11). We can clearly see that in Fig.(11)(a) linear coupling introduces power oscillation among the supermodes, whereas, in

Refer to caption
Figure 11: Comparison between the effect of linear and nonlinear (XPM) coupling during the evolution of the hybrid modes. Evolution of peak power (above) and total power evolution (below) in two supermodes under (a) only linear coupling and (b) only nonlinear coupling (XPM) (c) both in presence of linear and nonlinear XPM coupling; pumped at λ=1.567\lambda=1.567 μ\mum in SCR.

(b) XPM affects the coupling to the very beginning, after which pulse amplitude decays individually without any interaction. Total effect is shown in (c), which shows the dominant outcome of linear mode coupling between the two hybrid mode fields. As mentioned in Fujii2 , XPM is negligible for short pulse, and also because of choice of external parameters, we observe XPM induced phase-shift is quite negligible compare to the phase-shift due to linear mode coupling.

Appendix B Detailed form of Lagrangian and Rayleigh dissipation function

The individual form of LasL_{as},LsL_{s} for the Lagrangian and RasR_{as},RsR_{s} for the RDF, are as follows:

Las=i2(ψasψas,τψas,τψas)iD12(ψasψas,θψas,θψas)+D2as|ψas,θ|212|ψas|4+αas|ψas|2χ2(ψsψas+ψasψs),L_{as}=\frac{i}{2}(\psi_{as}\psi_{as,\tau}^{*}-\psi_{as,\tau}\psi_{as}^{*}){\color[rgb]{0,0,0}-\frac{iD_{1}}{2}(\psi_{as}\psi_{as,\theta}^{*}-\psi_{as,\theta}\psi_{as}^{*})}\\ +D_{2as}|\psi_{as,\theta}|^{2}-\frac{1}{2}|\psi_{as}|^{4}+\alpha_{as}|\psi_{as}|^{2}-\frac{\chi}{2}(\psi_{s}\psi_{as}^{*}+\psi_{as}\psi_{s}^{*}), (20)
Ls=i2(ψsψs,τψs,τψs)+iD12(ψsψs,θψs,θψs)+D2s|ψs,θ|212|ψs|4+αs|ψs|2χ2(ψasψs+ψsψas),L_{s}=\frac{i}{2}(\psi_{s}\psi_{s,\tau}^{*}-\psi_{s,\tau}\psi_{s}^{*}){\color[rgb]{0,0,0}+\frac{iD_{1}}{2}(\psi_{s}\psi_{s,\theta}^{*}-\psi_{s,\theta}\psi_{s}^{*})}\\ +D_{2s}|\psi_{s,\theta}|^{2}-\frac{1}{2}|\psi_{s}|^{4}+\alpha_{s}|\psi_{s}|^{2}-\frac{\chi}{2}(\psi_{as}\psi_{s}^{*}+\psi_{s}\psi_{as}^{*}), (21)
Ras=iSas(ψas,τψas,τ)+i(ψasψas,τψas,τψas),R_{as}=-iS_{as}(\psi_{as,\tau}^{*}-\psi_{as,\tau})+i(\psi_{as}\psi_{as,\tau}^{*}-\psi_{as,\tau}\psi_{as}^{*}), (22)
Rs=iSs(ψs,τψs,τ)+i(ψsψs,τψs,τψs),R_{s}=-iS_{s}(\psi_{s,\tau}^{*}-\psi_{s,\tau})+i(\psi_{s}\psi_{s,\tau}^{*}-\psi_{s,\tau}\psi_{s}^{*}), (23)

where, ψas,τ/s,τ=ψas,sτ\psi_{as,\tau/s,\tau}=\frac{\partial\psi_{\text{as,s}}}{\partial\tau} and ψas,θ/s,θ=ψas,sθ\psi_{as,\theta/s,\theta}=\frac{\partial\psi_{\text{as,s}}}{\partial\theta}. The following reduced forms of LgL_{g} and RgR_{g} are obtained after integrating the L(Las+Ls)\textit{L}(L_{as}+L_{s}) and R(Ras+Rs)\textit{R}(R_{as}+R_{s}) over azimuthal angle (θ\theta), mentioned in the main text. In order to execute the integration, we have to assume the width of the pulse in both modes remain same in the coupling term (containing χ\chi). Here, ϕ=ϕsϕas\phi=\phi_{s}-\phi_{as}.

Lg=2ηas(ϕasτD12+D2as(D12+ηas23)13ηas2+αas)2χηasΔscosϕcsch(Δas)+2ηs(ϕsτD12+D2s(D12+ηs23)13ηs2+αs)2χηsΔascosϕcsch(Δs),L_{g}=2\eta_{as}\left(\frac{\partial\phi_{as}}{\partial\tau}-{\color[rgb]{0,0,0}D_{1}^{2}}+D_{2as}\left(D_{1}^{2}+\frac{\eta_{as}^{2}}{3}\right)-\frac{1}{3}\eta_{as}^{2}+\alpha_{as}\right)\\ -2\chi\eta_{as}\Delta_{s}\cos\phi\operatorname{csch}\left(\Delta_{as}\right)\\ +2\eta_{s}\left(\frac{\partial\phi_{s}}{\partial\tau}-{\color[rgb]{0,0,0}D_{1}^{2}}+D_{2s}\left(D_{1}^{2}+\frac{\eta_{s}^{2}}{3}\right)-\frac{1}{3}\eta_{s}^{2}+\alpha_{s}\right)\\ -2\chi\eta_{s}\Delta_{as}\cos\phi\operatorname{csch}\left(\Delta_{s}\right), (24)
Rg=(4ηas2Sasπcosϕassech(Δas))ϕasτ(π2D1Sassinϕasηas2sech(Δas)tanh(Δas))ηasτ+(4ηs2Ssπcosϕssech(Δs))ϕsτ(π2D1Sssinϕsηs2sech(Δs)tanh(Δs))ηsτ,R_{g}=\Bigl{(}4\eta_{as}-2S_{as}\pi\cos\phi_{as}\operatorname{sech}\left(\Delta_{as}\right)\Bigr{)}\frac{\partial\phi_{as}}{\partial\tau}\\ -\left(\frac{\pi^{2}D_{1}S_{as}\sin\phi_{as}}{\eta_{as}^{2}}\operatorname{sech}\left(\Delta_{as}\right)\tanh\left(\Delta_{as}\right)\right)\frac{\partial\eta_{as}}{\partial\tau}\\ +\Bigl{(}4\eta_{s}-2S_{s}\pi\cos\phi_{s}\operatorname{sech}\left(\Delta_{s}\right)\Bigr{)}\frac{\partial\phi_{s}}{\partial\tau}\\ -\left(\frac{\pi^{2}D_{1}S_{s}\sin\phi_{s}}{\eta_{s}^{2}}\operatorname{sech}\left(\Delta_{s}\right)\tanh\left(\Delta_{s}\right)\right)\frac{\partial\eta_{s}}{\partial\tau}, (25)

Appendix C Stokes parameters and Jones vector for pump wavelength in SCR and WCR

In this section, we have mentioned the Stokes parameters (S0S_{0},S1S_{1},S2S_{2},S3S_{3}) and Jones vector (JJ) for two pump wavelengths, one in SCR (λ=1.567(\lambda=1.567 μ\mum), another in WCR (λ=1.614(\lambda=1.614 μ\mum). Stokes parameters are obtained

Table 2: Stokes parameters and Jones vector evolution with normalized slow time in strongly and weakly coupled region
Pump wavelength (λp)\lambda_{p}) Measured Quantity Normalized slow time (τ\tau)
τ=0\tau=0 τ=0.25\tau=0.25 τ=0.50\tau=0.50 τ=0.75\tau=0.75 τ=1.0\tau=1.0
1.567 μ\mum (SCR) S0S_{0} 4 2.356 1.340 0.767 0.446
S1S_{1} 0 -1.781 -0.754 0.086 -0.310
S2S_{2} 4 -0.246 0.152 0.746 -0.091
S3S_{3} 0 1.523 -1.098 0.156 0.307
J (1.4141.414e0.0i) (0.5361.438e-80.83i) (0.5411.023e-82.13i) (0.6530.583e11.84i) (0.2600.6148e-73.54i)
1.614 μ\mum (WCR) S0S_{0} 4 2.235 1.171 0.614 0.326
S1S_{1} 0 -0.066 -0.039 0.001 0.022
S2S_{2} 4 2.233 1.169 0.613 0.325
S3S_{3} 0 0.028 0.025 0.02 0.013
J (1.4141.414e0.0i) (1.0411.072e0.218i) (0.7520.779e.318i) (0.5540.553e.28i) (0.4170.389e.263i)

from coupled LLE (Eq.6-Eq.7) at θ=0\theta=0, and the corresponding Jones vector (JJ) are obtained with the help of Eq.(19). We have shown the evolution of these two quantities with slow time in Table. C.

References

  • (1) V. Brasch, E. Lucas, J. Jost, M. Geiselmann and T.  J.  Kippenberg, Light Sci. Appl. 6, e16202 (2017).

  • (2) B.  S.  Lee, M.  Zhang, F.  A.  S.  Barbosa, S.  A.  Miller, A.  Mohanty, R.  ST-. Gelais, and M.  Lipson, Opt. Express 25, 12110 (2017).
  • (3) T.  Hansson and S.  Wabnitz, Nanophotonics 5, (2016).
  • (4) T.  Herr, V.  Brasch, J.  D.  Jost, I.  Mirgorodskiy, G.  Lihachev, M.  L.  Gorodetsky, and T.  J.  Kippenberg, Phys. Rev. Lett. 113, 123901 (2014).
  • (5) M.  L.  Weichman, P.  B.  Changala, J.  Ye, Z.  Chen, M.  Yan and N.  Picqué, J. of Mol. Spectrosc. 355, 66 (2019).
  • (6) A.  Fülöp, M.  Mazur, A.  L-.  Riesgo, T.  A.  Eriksson, P-. H.  Wang, Y.  Xuan, D.  E.  Leaird, M.  Qi, P.  A.  Andrekson, A.  M.  Weiner and V.  T-.  Company, Opt. Express 25, 26678 (2017).
  • (7) S.  B.  Papp, K.  Beha, P.  Del’Haye, F.  Quinlan, H.  Lee, K.  J.  Vahala, and S.  A.  Diddams, Optica 1, 10, (2014).
  • (8) P.  Trocha, M.  Karpov, D.  Ganin, M.  H.  P.  Pfeiffer, A.  Kordts, S.  Wolf, J.  Krockenberger,P.  Marin-Palomo, C.  Weimann, S.  Randel, W.  Freude, T.  J.  Kippenberg and C.  Koos, Science 359, 6378 (2018).
  • (9) H.  Guo, M.  Karpov, E.  Lucas, A.  Kordts, M.  H.  P.  Pfeiffer, V.  Brasch, G.  Lihachev, V.  E.  Lobanov, M.  L.  Gorodetsky and T.  J.  Kippenberg, Nat. Phys. 13, 94 (2017).
  • (10) C.  C.  Lam, P.  T.  Leung, and K.  Young, J. Opt. Soc. Am. B 9, 1585 (1992).
  • (11) R. Haldar, A. Roy, P. Mondal, V. Mishra, and S.  K. Varshney, Phys. Rev. A 99, 033848 (2019).
  • (12) A. Roy, R. Haldar and S. K.  Varshney, J. Lights. Technol. 36, 5807 (2018).
  • (13) S.  Fujii, Y.  Okabe, R.  Suzuki, T.  Kato, A.  Hori, Y.  Honda and T.  Tanabe, IEEE Photonics. J. 10, 4501511 (2018).
  • (14) 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, Optica 1, 2334, (2014).
  • (15) S. Fujii, A. Hori, T. Kato, R. Suzuki, Y. Okabe, W. Yoshiki, A. C. Jinnai and T. Tanabe, Opt. Express 25, 28969 (2017).
  • (16) W.  Yoshiki, A.  C. Jinnai, T.  Tetsumoto, and T.  Tanabe, Opt. Express 23, 30852 (2015).
  • (17) M. Saha, S. Roy, S. K. Varshney, Phys. Rev. A, 101 033826 (2020).
  • (18) R. Suzuki, S. Fujii, A. Hori, T. Tanabe, IEEE Photonics. J. 11, 6100511 (2019).
  • (19) T. Hansson, M. Bernard and S. Wabnitz, J. Opt. Soc. Am. B 35, 835 (2018).
  • (20) G. D’Aguanno and C. R. Menyuk, Eur. Phys. J. D 71, 1 (2017).
  • (21) E.  Lucas, M.  Karpov, H.  Guo, M. L.  Gorodetsky and T. J.  Kippenberg, Nat. Commun. 8, 736 (2017).
  • (22) J.  Wiersig, Phys. Rev. Lett 97, 253901 (2006).
  • (23) M.  Soltani, A.  Matsko and L.  Maleki, Laser Photonics Rev. 10, 158 (2016).
  • (24) S. Roy and S. K. Bhadra, J. Lights. Technol. 26, 2301 (2008).
  • (25) M. Saha, S. Roy and S. K. Varshney, Phys. Rev. E 100, 022201 (2019).
  • (26) A.  Sahoo and S. Roy, Phys. Rev. A, 100 053814 (2019).
  • (27) S.  Kim, K.  Han, C.  Wang, J.  A.  Jaramillo-Villegas, X.  Xue, C.  Bao, Y.  Xuan, D.  E.  Leaird, A.  M.  Weiner and M.  Qi, Nat. Commun. 8, 372 (2017).
  • (28) K. Okamoto, Fundamentals of Optical Waveguides, (Academic press, 2006).
  • (29) P. Grelu, Nonlinear Optical Cavity Dynamics: From Microresonators to Fiber Lasers, 1st ed. (Wiley-VCH, 2016).
  • (30) L. A. Lugiato and R.  Lefever, Phys. Rev. Lett. 58, 2209 (1987).
  • (31) D. Anderson, Phys. Rev. A 27, 3135 (1983).
  • (32) H.  A.  Haus and W.  Huang, Proceedings Of The IEEE 79, 1505 (1991).
  • (33) E. Averlant, M. Tlidi, K. Panajotov and L. Weicker, Opt. Lett. 42, 2750 (2017).
  • (34) R. A. Chipman, W. -S. T. Lam, G. Young, Polarized Light and Optical Systems, 2nd ed. (CRC-Press, 2018).
  • (35) G. P. Agrawal, Nonlinear Fiber Optics, 5th ed. (Academic, New York, 2012)
  • (36) J. N. Damask, Polarization Optics in Telecommunications, (Springer, New York, 2005 )