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

In-phase, out-of-phase and T/4 synchronization of square waves in delay-coupled non-identical optoelectronic oscillators

Open Access Open Access

Abstract

We model two non-identical delay-line optoelectronic oscillators mutually coupled through delayed cross-feedback. The system can generate multi-stable nanosecond periodic square-wave solutions which arise through a Hopf instability. We show that for suitable ratios between self and cross delay times, the two oscillators generate square waves with different amplitude but synchronized in phase, out of phase or with a dephasing of a quarter of the period. We also show that the synchronization is robust to small mismatches in the delay times.

© 2015 Optical Society of America

1. Introduction

Time delays appear naturally in a variety of physical, biological, chemical and technological systems, and are a source of oscillatory instabilities [1]. When the delay is large compared to the other time scales of the system, square-waves are the dominant solutions for specific values of the parameters. Motivated by applications such as optical clocks [2] or optical sensing [3], the generation of stable pulses with square-wave form has been investigated in optical and optoelectronic systems during the past few years [2–10].

The possibility to generate square-wave pulses of controllable period has been discussed in the context of the Ikeda differential equation [11–13], which presents a nonlinear feedback term proportional to cos[x(tτ) + Φ], where τ is the delay time and Φ a constant offset phase which fixes the operating point, namely, depending on the value of Φ the effect of the feedback term is to damp small perturbations (negative feedback) or to amplify them (positive feedback). The Ikeda equation was originally introduced to describe the dynamics of the transmitted light from a nonlinear absorbing medium subject to delayed optical feedback [14] and has been later generalized to describe other systems with delayed feedback [15, 16], among them, optoelectronic oscillators (OEOs). While several OEO configurations have been considered [17, 19], including those based on whispering gallery modes [18] and multicore fibers [20], here we focus on delay-line OEOs [21] which can be modeled by an Ikeda-like equation with feedback proportional to cos2[x(tτ) + Φ] [22]. Feedback is negative for offset phases Φ [0, ∈ π/2] and positive for Φ ∈ [−π/2, 0]. Square waves originated with negative feedback have a symmetric duty cycle, namely the duty cycle is half the period. In contrast, for positive feedback square waves have an asymmetric duty cycle. Generation of stable square waves with symmetric and asymmetric duty cycles in OEOs has been reported in Refs. [23–25]. Stable square waves have also been found in other optical systems such as edge-emitting diode lasers [4–6], vertical cavity surface-emitting lasers [7,8,26], ring lasers [9] or mode-locked fiber lasers [10,27].

Coupling with delay two oscillators, each with an intrinsic long delay so that it generates square-waves, leads to interesting dynamics arising from the interplay between the intrinsic and the coupling delay times. In particular, two mutually coupled identical OEOs with negative feedback can generate stable in-phase and out-of-phase synchronized square waves with symmetric duty cycle when the ratio between the delay times fulfills suitable conditions [28]. For positive feedback in- and out-of-phase synchronization is also obtained but in this case the duty cycle is not symmetric [29]. In general, the square-waves generated so far with optical and optoelectronic systems have been obtained synchronized in phase or out of phase. In this work we show that two mutually coupled non-identical OEOs, besides in-phase and out-of-phase, can generate square-waves synchronized at a quarter of the period (T/4).

Synchronization at a quarter of the period has been observed in some animal gaits of quadrupeds as walk and jump [30], in the limb coordination in crustacean swimming [31], between the oscillations of pedestrian walking on a bridge and the structural oscillation [32], in predator-prey cycles [33–35], in human cortical sources [36], neural networks [37] and in solar transverse oscillations [38]. Nevertheless, in none of these studies the oscillations have a square-wave shape. Here by exploring the dynamical regimes that can arise in a prototypical model for mutually delay-coupled OEOs we show that stable square-wave pulses synchronized at a quarter of the period do exist in a broad parameter region. The key point to obtain such T/4 solutions is that the two OEO operate with different offset phases, in such a way that the feedback is negative in one and positive in the other, namely the feedback is mixed. It should also be emphasized that these T/4 solutions are in the ns time scale, orders of magnitude faster than the above systems.

The outline of the article is as follows. We describe the model in Sec. 2. In Sec. 3 we study the Hopf instabilities of the steady state, in Sec. 4 the in- and out-of-phase square waves with symmetric duty cycle occurring when both OEOs have negative feedback, in Sec. 5 the periodic solutions with asymmetric duty cycle arising when both OEOs have positive feedback, in Sec. 6 the square waves synchronized at a quarter of the period, and in Sec. 7 the effect of a small mismatch in the delay times on the degradation of the T/4 solutions. Sec. 8 concludes the paper.

2. System description

We consider two OEOs, similar to those of [39], mutually coupled with delay, both fed by half of the light from a cw semiconductor laser (LD) emitting with intensity P as illustrated in Fig. 1. Each OEO consists of a Mach-Zehnder interferometer (MZI), an optical delay line, a photodiode (PD) with sensitivity S and an amplifier. The variables related to loop i are identified with subindex i, i = 1, 2. The optical power after MZIi is split into two parts. A fraction αii is delayed Tii within OEOi (self-feedback) and a fraction αij with ji couples OEOi with OEOj with a delay time Tij (cross-feedback). A photodiode (PD) detects the intensity resulting from the combination of the self- and cross-feedback optical beams and produces an electrical signal, which is amplified and band-pass filtered, and finally used to drive the Mach-Zehnder AC electrode. Assuming no reflections in the optical paths, we can take the optical electric field as a scalar. We also consider that the voltage Vi applied to MZIi has a DC and a RF component, VBi and VRFi(t) respectively, and introduce the dimensionless variables xi(t)=πVRFi(t)/2Vπi and Φi=πVBi/2Vπi, being Vπ the voltage required for a change of π in the phase. Introducing two additional variables, yi(t)=t0txi(t)dt, the dynamics of the system is ruled by [28],

τix˙i(t)=xi(t)θi1yi(t)+P{γii2cos2[xi(tTii)+Φi]+γji2cos2[xj(tTji)+Φj]+2γiiγjicos[xj(tTji)+Φj]cos[xi(tTii)+Φi]×cos[xi(tTii)xj(tTji)+ΦiΦj+(1)iΦ0]}y˙i(t)=xi(t),
where i, j = 1,2, ji, θi and τi are the band-pass filter low and high cut-off times, P is the pump parameter, Φ0 is the phase difference between the non-modulated arms of the two MZIs and γijαij are effective feedback strengths. For simplicity we consider θ1 = θ2 = θ, τ = τ2 = τ and T11 = T22 = Tf and define Tc = (T12 + T21)/2. This system has a steady state solution,
xist=0yist=θP{γii2cos2Φi+γii2cos2Φj+2γiiγjicosΦjcosΦicos[Φ1Φ2Φ0]},ji.

 figure: Fig. 1

Fig. 1 Diagram of the system modeled integrated by two mutually delay-coupled OEOs. Each OEO consists of a Mach-Zehnder interferometer (MZI), a fiber loop with delay time Tf, a photodiode (PD) and a RF amplifier (G) whose output modulates an arm of the MZI. OEOs are fed by a laser diode (LD) whose output is split in two parts by a 50/50 fiber splitter. The OEOs are mutually coupled with cross-feedback delay time Tc.

Download Full Size | PDF

Introducing Yi(t)=[yi(t)yist]/Tc and scaling the time with Tc, s = t/Tc, we get

εxi(s)=xi(s)δYi(s)+Pγii2{cos2[xi(ss0)+Φi]cos2Φi}+Pγji2{cos2[xj(s1)+Φj]cos2Φj}+2Pγiiγji{cos[xi(ss0)+Φi]cos[xj(s1)+Φj]×cos[xi(ss0)+xj(s1)+ΦjΦi+(1)iΦ0]cosΦicosΦjcos(Φ1Φ2Φ0)}Yi(s)=xi(s),
where prime means differentiation with respect to s and
s0=Tf/Tc,ε=τTc1,δ=Tcθ1.

3. Oscillatory instabilities of the steady state

Equations (3) admit periodic square-wave solutions in the limit of large delays which are born as sinusoidal solutions from a Hopf bifurcation of the steady state xist=Yist=0 which takes place increasing P. Thus we first analyze the steady state stability by considering the linearized equations for the small perturbations, Ui(s)=xi(s)xist and Vi(s)=Yi(s)Yist,

εUi(s)=Ui(s)δViP[FiUi(ss0)+KiUj(s1)]Vi(s)=Ui(s),
where the effective self-feedback and coupling coefficients are given by
Fi=γii2sin2Φi+2γiiγjicosΦjsin[2ΦiΦj+(1)iΦ0],Ki=γji2sin2Φj+2γiiγjicosΦisin[2ΦjΦi(1)iΦ0].

Equations (5) admit solutions of exponential form, Ui(s) = ui exp[(λ + )s] and, since Vi is the integral of Ui, Vi(s) = ui exp[(λ + )s]/(λ + ). Replacing Ui and Vi in Eqs. (5) leads to

0=[1+ε(λ+iω)+δ(λ+iω)1+PFie(λ+iω)s0]ui+PKie(λ+iω)uj,

The condition for non-trivial solutions is that the determinant vanishes, namely

0=[1+ε(λ+iω)+δ(λ+iω)1+PF1e(λ+iω)s0]×[1+ε(λ+iω)+δ(λ+iω)1+PF2e(λ+iω)s0]P2K1K2e2(λ+iω).

The steady state becomes unstable at the bifurcation point P = Pc where λ = 0. Splitting Eq. (8) into real and imaginary parts we obtain for λ = 0:

0=1+Pc(F1+F2)[cos(ωs0)+(εωδω)1sin(ωs0)]+Pc2[F1F2cos(2ωs0)K1K2cos(2ω)](εωδω1)2,
0=Pc(F1+F2)[(εωδω1)cos(ωs0)sin(ωs0)]Pc2[F1F2sin(2ωs0)K1K2sin(2ω)]+2(εωδω1).

The solutions of Eq. (9) can be seen as the zeros of a second order polynomial for Pc whose coefficients depend on the system parameters and the perturbation frequency ω. We first consider the case ε = δ = 0. The value of the polynomial is 1 for Pc = 0, thus for Pc small Eq. (9) has no solutions meaning that the steady state is stable. For given Fi, Ki and s0, the first instabilities are those such that the linear coefficient, (F1 +F2)cos(ωs0), takes the most negative value, namely

ωs0=kπ,
being k′ an integer, odd for F1 + F2 > 0 and even for F1 + F2 < 0. Then Eq. (10) leads to K1K2 sin(2ω) = 0, namely
ω=kπ/2.

For Eqs. (12) and (11) to be mutually compatible it is required that,

s0=2kk.

Thus the instabilities with lower threshold take place for s0 rational. The period of the emerging oscillatory solutions is given by

T=2(1s0)2kk=2s0k=4k.

The Hopf instability threshold, Pc, can be obtained substituting Eqs. (12) and (11) in Eq. (9): else the smaller positive solution for Pc is

1+Pc(F1+F2)(1)k+Pc2(F1F2K1K2)(1)k)=0.

If (F1 F2) = K1K2(−1)k than

Pc=(1)kF1+F2,
else the smaller positive solution for Pc is
Pc=(F1+F2)(1)k(F1F2)2+4K1K2(1)k2[F1F2K1K2(1)k].

In order Eq. (17) to be real it is required that (F1F2)2 + 4K1K2(−1)k > 0. Since Pc depends only on the parity of k′ and k many bifurcations have the same threshold.

Introducing Eqs. (12) and (11) in Eq. (7) leads to a ratio for the perturbation amplitudes:

u2=Qu1,Q=(i)k1+PcF1(1)kPcK1.

For k even, Q is real and there are two kinds of solutions depending on the sign of Q. For Q > 0 u1 and u2 have the same sign and the instability leads to oscillatory solutions in which x1 and x2 are in phase (although the amplitude may not be the same) while for Q < 0 the instability leads to out-of-phase oscillations. For k odd, Q is a purely imaginary number. Considering xi = Ui + c.c. this leads to x1 and x2 being dephased by T/4.

In Fig. 2 we plot Pc obtained from Eqs. (16) and (17) for γij = γ = 0.5 and different parities of k and k′. Provided γij are identical, the same results are obtained for other values of γ rescaling Pc with γ2. Pink regions correspond to offset phases for which Pc > 5, while white and grey regions correspond to unphysical solutions for Pc (negative or complex values respectively). For Φ1 ≃ Φ2, along the diagonal, only instabilities associated to k even can take place [panels (a) and (b)]. Thus, as pump is increased, the steady state becomes unstable to either in-phase or out-of-phase oscillations. On the contrary, for Φ1 = −Φ2, along the anti-diagonal, the only possible instabilities are those with k odd [panels (c) and (d)] leading to periodic oscillations dephased by T/4. In fact, there is a broad parameter region where the lower threshold corresponds to k-odd instabilities leading to T/4 dephased oscillations.

 figure: Fig. 2

Fig. 2 Pc as given by Eqs. (16)(17) for γij = 0.5 and Φ0 = 0. (a) k′ odd and k even, (b) k′ even and k even, (c) k′ odd and k odd, (d) k′ even and k odd. Parameter regions in which Pc is negative or imaginary are plotted in white and grey, respectively.

Download Full Size | PDF

For non-identical γij regions in (Φ12) parameter space get distorted as illustrated in Fig. 3. Nevertheless there are regions where k-even instabilities dominate, such as Φ1 ≈ Φ2, leading to in-phase and out-of-phase solutions, and regions where k-odd instabilities leading to T/4 dephased solutions are the first to take place increasing the pump. In general, physical solutions for Pc with k′ odd and k even [panel (a) in Figs. 2 and 3] exist for offset phases such that F1 + F2 > 0 (negative feedback), solutions with k′ even and k [panel (b)] odd exist for offset phases such that F1 + F2 < 0 (positive feedback), and solutions with k odd [panels (c) and (d)] exist for offset phases such that K1K2 < 0 (mixed feedback).

 figure: Fig. 3

Fig. 3 Value of Pc as in Fig. 2 for γ11 = 0.5, γ22 = 0.3, γ12 = 0.2 and γ21 = 0.4.

Download Full Size | PDF

We note that for Φ0 = 0, changing the sign of Φ1 and Φ2 changes the sign of Fi and Ki so that Eqs. (16) and (17) are the same replacing k′ by k′ + 1. As a consequence, the instabilities associated to k′ odd occur for offset phases which are the specular image of those for k′ even [compare panel (a) with (b) and (c) with (d) in Figs. 2 and 3]. This symmetry is broken for Φ0 ≠ 0.

In the general case of ε ≠ 0, δ ≠ 0 and arbitrary s0 we can determine Pc and ω solving numerically Eqs. (9) and (10). From Eq. (7) we can also obtain a theoretical prediction for Q = u2/u1, which is a complex number whose modulus,

|Q|=[1+(εωδω1)2+Pc2F12+2PcF1[cos(ωs0)(εωδω1)sin(ωs0)]1/2Pc|K1|,
gives the ratio between the oscillation amplitudes, and whose phase gives the phase difference between x1 and x2,
φ2φ1=arctansin(ω)+(εωδω1)cos(ω)+PcF1sin[ω(1s0)]cos(ω)(εωδω1)sin(ω)+PcF1cos[ω(1s0)].

In the next sections we analyze the in- and out-of-phase solutions arising for negative and positive feedback, and the T/4 dephased solutions for mixed feedback.

4. Onset of periodic square waves for negative feedback

As discussed in Ref. [28] for identical γij, Φ0 = 0 and Φ1 = Φ2 = Φ ∈ [0,π/2], increasing the pump the steady state becomes unstable to in-phase or out-of-phase periodic oscillations. The instability is supercritical and, as the pump is further increased, the oscillations soon become square-shape with a duty cycle half of the period. Owing to parameter symmetry the oscillations in x1 and x2 have the same amplitude. Here we show that square waves with symmetric duty cycle can also be found for non-identical γij or offset phases (although x1 and x2 have different amplitudes) in a broad parameter region associated to negative feedback in both OEOs. More precisely they arise for parameter values such that the steady state instabilities associated to k′ odd and k even dominate [compare panel (a) in Figs. 2 or 3 with the other panels].

To illustrate this we consider Φ1 = 0.2π, Φ2 = 0.3π and γij = 0.5. Fig. 4 shows all the Hopf bifurcations in the range ω < 10π and 0 < Pc < 2 obtained solving numerically Eqs. (9) and (10). For these parameters, Eq. (17) has physical solutions only for k′ odd and k even (See Fig. 2). Thus, the bifurcations with lower threshold occur at s0 = 2k′/k = k′/n (n = k/2 integer), namely at rational fractions s0 = q′/q (irreducible) with odd numerator. At these values of s0 a set of bifurcation lines with k′ = jq′ and k = 2jq (j odd) are born simultaneously [Fig. 4(a)]. j = 1 corresponds to the fundamental solution with ω = while j = 3,5,… correspond to harmonics with ω = jqπ. Moving away from s0 = q′/q the degeneracy is broken and Pc has a parabolic dependence with s0q′/q, the parabola being narrower for higher harmonics. For q odd, k = 2jq is even but not multiple of 4 (singly even) and Q > 0, so that all the harmonics are in phase (plotted in red). Conversely, for q even, k = 2jq is multiple of 4 (doubly even) and Q < 0, which leads to out-of-phase oscillations (black lines).

 figure: Fig. 4

Fig. 4 Hopf bifurcations with ω < 10π for Φ1 = 0.2π, Φ2 = 0.3π, Φ0 = 0, γij = 0.5, ε = 4.17 × 10−4 and δ = 1.2 × 10−2. In (b) lines are plotted only in the range where Pc < 2. Red and black lines correspond to in- and out-of-phase oscillations, respectively.

Download Full Size | PDF

The previous results have been obtained from the linear stability analysis of the steady state. To analyze to which extend this results are valid far away from the bifurcation point we resource to numerical integration of the full nonlinear dynamical equations (3). Figure 5 shows the fundamental out-of-phase solution for Φ1 = 0.2π and Φ2 = 0.25π and s0 = 3/4, so that k′ = 3 and k = 8. For the parameters of the figure the steady state becomes unstable at Pc = 2.1167, giving rise to out-of-phase quasi-sinusoidal oscillations [Fig. 5(a)] with a period T = 1/2 (20ns) as predicted by Eq. (14). While the two OEOs are synchronized in frequency, amplitudes differ because each oscillator receives a different feedback strength. The ratio Q between the amplitudes of the oscillations in x1 and x2 obtained from the numerical integration of (3), Q = −0.2529, is in good agreement with the prediction of the linear stability analysis (19), Q = 0−.2521, and also with the approximation (18), Q = −0.2519. Increasing the pump the shape of the solution becomes less and less sinusoidal approaching a square-wave [Figs. 5(b) and 5(c)]. Nevertheless the solutions remain synchronized out of phase and the ratio between the amplitudes Q remains practically constant. Similar good agreements are found for in-phase solutions arising at delay times fulfilling the ratio s0 = q′/q with q′ and q odd.

 figure: Fig. 5

Fig. 5 Out-of-phase oscillations with symmetric duty cycle for Φ1 = 0.2π, Φ2 = 0.25π, Φ0 = 0, γ11 = 0.5, γ22 = 0.3, γ12 = 0.2, γ21 = 0.4, Tf = 30ns, Tc = 40ns, ε = 6.25 × 10−4, δ = 8 × 10−3 and P = 2.117 (a), P = 2.12 (b), and P = 2.3 (c). Black and red lines correspond to x1 and x2 respectively.

Download Full Size | PDF

5. Onset of periodic oscillations for positive feedback

For Φ0 = 0 and Φ1 = Φ2 = Φ ∈ [−π/2,0] square waves typically have an asymmetric duty cycle and are born subcritically [29]. The degree of asymmetry depends on the offset phase, so that the duty cycle is symmetric (and square waves are supercritical) for Φ = −π/4 and becomes progressively more asymmetric as the offset phase is changed away form −π/4. For delay times whose ratio is rational, s0 = q′/q irreducible, for q′ odd the system displays in- phase square waves while for q′ even both in-phase and out-of-phase asymmetric square waves coexist. The system can also display periodic oscillations with a period much longer than the delay times, typically in the microsecond regime. These oscillations are always in phase, born supercritically and do not have a square-wave shape. We show here that a similar scenario occurs for non-identical offset phases and parameter values for which instabilities associated to k′ and k both even dominate [compare (b) in Figs. 2 or 3 with the other panels].

To illustrate this we consider Φ1 = 0.2π, Φ2 = 0.3π and γij = 0.5. Figure 6 shows all the Hopf bifurcations in the range ω < 10π and 0 < Pc < 1.15 from numerical solution of Eqs. (9) and (10). For these parameters, Eq. (17) has physical solutions only for k′ odd and k even (See Fig. 2). A set of Hopf bifurcation lines are born simultaneously at rational values for the delay ratio s0 = q′/q, irreducible. Different Hopf lines correspond to different values of k′ = jq′ and k = 2jq. For q′ odd, j must even, thus k is doubly even and Q > 0, which leads to in-phase oscillations. For q′ even any j is allowed (q′/q irreducibility requires q odd which is not restrictive since k = 2jq is always even). The fundamental solution j = 1 and the odd harmonics have a singly even k and Q < 0 and thus are out of phase, while the even harmonics are in phase. As before degeneracy is broken for s0 different from q′/q and Hopf lines have a parabolic shape which is narrower for higher harmonics.

 figure: Fig. 6

Fig. 6 Hopf bifurcations for Φ1 = −0.2π and Φ2 = −0.3π. Other parameters as in Fig. 4. The green line corresponds to microsecond oscillations.

Download Full Size | PDF

For identical γij and Φ0 = 0 square waves have a symmetric duty cycle and arise supercritically if Φ12 = −π/2, as the fundamental out-of-phase solution for Tf = 40ns and Tc = 60ns (s0 = 2/3) shown in Fig. 7. Here k′ = 2, k = 6 and the solution has a period T = 2/3 (40ns). For other offset phases oscillations are born subcritically and have an asymmetric duty cycle, as shown in Fig. 8 for the same delay times but with non-identical γij, Φ1 = −0.25π, and Φ2 = −0.15π. For the parameters of Fig. 8, Pc = 1.6334. As pump is increased above threshold the periodic solution becomes more square-wave shaped while the duty cycle asymmetry decreases [Figs. 8(c) and 8(d)]. Below threshold, decreasing pump the duty cycle asymmetry increases until the length of the shorter plateau vanishes [Fig. 8(a), with a pulse width 0.044T]. The duty cycle asymmetry also increases moving the offset phases away from Φ12 = −π/2. The minimum duty cycle is set by the transition time between the square-wave plateaus, which is determined by τ; thus, it is of order ε in dimensionless time. Other than that, the duty cycle can be conveniently tuned over the whole period, while keeping the period constant, by changing the pump or the DC voltages applied to the MZIs as shown in Ref. [29] for Φ1 = Φ2.

 figure: Fig. 7

Fig. 7 Fundamental out-of-phase solution for Φ1 = −0.1π, Φ2 = −0.4π, Φ0 = 0, γii = 0.5, Tf = 40ns, Tc = 60ns, ε = 4.17 × 10−4, δ = 1.2 × 10−2, P = 1.702 (a), P = 1.72 (b), and P = 1.8 (c). Pc = 1.7013.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Fundamental out-of-phase solution for Φ1 = −0.25π, Φ2 = −0.15π, Tf = 40ns, Tc = 60ns, ε = 4.17×10−4, δ = 1.2×10−2, P = 1.616 (a), P = 1.6271 (just below threshold) (b), P = 1.7 (c), and P = 2 (d). Other parameters as in Fig. 5.

Download Full Size | PDF

We now focus on the microsecond oscillations associated to j = 0 (k′ = k = 0). Neglecting ε and δ, according to Eq. (11), ω = 0. For finite ε and δ, ω is slightly shifted from Eq. (11):

ωk=kπs0+αk,
where αk is small compared to kπs0, and can be neglected, except for k′ = 0. In this case, expanding Eq. (10) to first order in α0 one has
0=Pc(F1+F2)(εα0δα0α0s0)2Pc2α0(F1F2s0K1K2)+2(εα0δα0),
which leads to
α02=[Pc(F1+F2)+2]δPc(F1+F2)(εs0)2Pc2(F1F2s0K1K2)+2ε.

Therefore ω0 = α0δ1/2 in the MHz regime typically. As shown in Fig. 6 these solutions have a threshold value which is practically independent of s0 (green line) and which takes place at a value slightly larger than the minimum threshold for square waves [29]. Low frequency Hopf instabilities are also found in a single OEO with positive feedback [23, 40] and can be used for pure microwave generation [21,41,42]. For two coupled OEOs Fig. 9 shows the shape of a microsecond solution with asymmetric duty cycle for different pump values. Increasing the pump the asymmetry of the duty cycle increases and the solution shows a tilted plateau, however it never takes a full square wave shape, rather becomes chaotic. The period of these solutions increases with the pump [29] as clearly seen in the figure. For identical γij and offset phases, microsecond solutions are in phase and have exactly the same amplitude. When offset phases differ, as in Fig. 9, they have a small dephasing as shown in panel (d). The dephasing is so small that at the scale of the other panels the traces for x1 and x2 overlap.

 figure: Fig. 9

Fig. 9 Microsecond solution obtained for Φ1 = −0.15π, Φ2 = −0.1π, and P = 1.444768 (a), P = 1.4538 (b), and P = 1.454 (c). Panel (d) shows a zoom of c) close to the maximum. Other parameters as in Fig. 7. The threshold is Pc = 1.4392 and at threshold ω0 = 0.1215. Time traces for x1 and x2 overlap in panels (a)–(c).

Download Full Size | PDF

6. Onset of T/4 dephased periodic square waves

Here we consider the parameter regions for which the instabilities of the steady state associated to k odd dominate. As shown in Figs. 2 and 3, in the (Φ1, Φ2) parameter space, this typicality occurs in the second and fourth quadrant where offset phases have opposite signs.

To illustrate the scenario, let us consider γij = 0.5, Φ1 = −0.25π and Φ2 = 0.15π, parameters for which increasing the pump the first instabilities are those associated to k odd and k′ even (Pc = 1.1188) [Fig. 2(d)] followed by those with k and k′ odd at Pc = 1.2526 [Fig. 2(c)]. Instabilities with k even are not relevant (either are not allowed or have a very large threshold [Figs. 2(a), 2(b)]). Figure 10 shows the Hopf bifurcations in the range ω < 8π and 0 < Pc < 1.8. In addition to Pc and ω obtained solving Eqs. (9) and (10) numerically, we have plotted the phase difference φ2φ1 from Eq. (20). Families of Hopf bifurcation lines are born at rational values of the delay ratio s0 = q′/q, irreducible. The members of a family can be identified by j, such that 2k′ = jq′ and k = jq. Since k is odd, q and j must be odd; and since 2k′ is even, q′ must be even. Thus Hopf lines are born at even-odd rational values for s0 and only odd harmonics are allowed. The stability analysis also predicts that instabilities arising at q′ doubly even (k′ even) have a lower threshold than those at q′ singly even (k′ odd) as shown in Fig. 10(a). The amplitude ratio and dephasing between x1 and x2 are given by the amplitude and phase of Q. At the onset of the Hopf lines, from Eq. (18)Q is purely imaginary, so there are two kinds of solutions: those where φ2φ1 = π/2 for which x2 is delayed a quarter of the period with respect to x1 (dephasing T/4), and those where φ2φ1 = −π/2 for which x2 advances x1 by a quarter of the period (dephasing −T/4). For the parameters considered in Fig. 10, |Q| = 0.448 for q′ doubly even while |Q| = 2.621 for q′ singly even. In both cases [1 + PcF1 (−1)k′]/(PcK1) > 0, thus the phase of Q is π/2 for k (mod 4) = 3 and −π/2 for k (mod 4) = 1. As a consequence, fundamental solutions born at s0 with q = 3,7,11,… are dephased T/4 while those born at s0 with q = 1,5,9,… are dephased −T/4, and successive harmonics born at a given s0 have alternatively positive and negative detunings as shown in Fig. 10. According to Eq. (20), changing s0 with respect to the rational q′/q leads to a change in the dephasing as shown in Fig. 10.

 figure: Fig. 10

Fig. 10 Hopf lines for Φ1 = −0.25π and Φ2 = 0.15π. Other parameters as in Fig. 4. Yellow and blue lines correspond to bifurcations leading to synchronized solutions dephased +T/4 and −T/4, respectively.

Download Full Size | PDF

T/4 square waves have a symmetric duty cycle and multiple stable harmonics can coexist, as illustrated in Fig. 11 for s0 = 4/3 (q′ doubly even with lower threshold) and for s0 = 2/3 (q′ singly even with larger Pc). In agreement with the stability analysis, for s0 = 4/3 the amplitude of x2 is smaller than that of x1 while the opposite occurs for s0 = 2/3. Also, while P is the same, owing to the lower threshold, the overall amplitude of the square waves for s0 = 4/3 is larger. Nevertheless, in both cases the fundamental solution corresponds to k = 3, and according to Eq. (14) has a period T = 4/3 (80ns since Tc = 60ns). First and second harmonics correspond respectively to j = 3 (k = 9) and j = 5 (k = 15) with period T = 4/9 (26.6ns) and 4/15 (16ns). For the fundamental and second harmonic x2 is delayed T/4 with respect to x1, while for the first x1 is delayed T/4 with respect to x2 (dephasing −T/4). Higher order harmonics are also found provided the duration of the plateau is longer than the transition time which, as discussed before, depends on τ. Figures 11(d) and (h) show the 30th harmonic, j = 61 (k = 183) with a period T = 4/183 (1.31ns). For τ = 25ps as used in Fig. 11 harmonics above 40 are unstable. However, for smaller τ, one can generate T/4 synchronized square waves with a subnanosecond period as shown in Fig. 12.

 figure: Fig. 11

Fig. 11 Coexistence of T/4 square waves for the parameters of Fig. 10, P = 1.3, and Tc = 60ns. Fundamental, 1st, 2nd and 30th harmonics for Tf = 80ns (s0 = 4/3) are shown in (a)–(d) while (e)–(f) show the corresponding ones for Tf = 40ns (s0 = 2/3). Black and red lines correspond to x1 and x2 respectively.

Download Full Size | PDF

 figure: Fig. 12

Fig. 12 Higher harmonics for the parameters of Fig. 11 but τ = 10ns and s0 = 4/3: (a) 40th, T = 4/243 (0.99ns), (b) 60th, T = 4/363 (0.66ns) and (c) 80th, T = 4/483 (0.50ns).

Download Full Size | PDF

T/4 oscillatory solutions are born supercritically with a sinusoidal shape as shown in Fig. 13(a) for the fundamental solution born at s0 = 4/3. Increasing the pump, the solution soon becomes square-wave shaped as it can be seen in Figs. 13(b) and Figs. 13(c).

 figure: Fig. 13

Fig. 13 Square-waves dephased T/4 for Tc = 60ns, Tf = 80ns, P = 1.119 (a), P = 1.13 (b), and P = 1.25 (c). Other parameters as in Fig. 11.

Download Full Size | PDF

7. Robustness of T/4 dephased square waves to delay time mismatch

We have shown that the Hopf lines leading to different kinds of synchronized square waves originate at suitable rational values for the ratio between the delay times, thus from an practical point of view, it is important to assess the robustness of the synchronized square waves to small mismatches in the delay times with respect to the ideal ratio. For in- and out-of-phase square waves arising for identical offset phases, we have shown that while synchronization is robust to mismatches up to a 5%, solutions with asymmetric duty cycle born for positive feedback are more robust than those with symmetric duty cycle arising for negative feedback, and that the larger the period of the solution the higher the robustness of the synchronization [28,29]. In this section we focus on the effect of a small mismatch in the delay times on the period, the shape and the synchronization of the T/4 dephased square waves.

As an example, we consider the fundamental and the first harmonic square-waves originated from the Hopf lines that in Fig. 10 exhibit a minimum at s0 = 4/3 (plotted in yellow and blue respectively). The left panels on Figs. 14 and 15 show the time traces obtained integrating numerically Eq. (3) for P = 1.25, keeping fixed Tc and changing Tf. Figure 14 has been obtained using the fundamental square wave (T = 4/3) as initial condition while the first harmonic (T = 4/9) has been used for Fig. 15. With 1% mismatch [Figs. 14(b) and Figs. 15(c)] the solutions exhibit longer transition times between the two the plateaus while the plateau lengths are reduced. Furthermore the slope of the positive jump (transition from the negative to the positive plateau) is different than the slope of the negative jump (transition from the positive to the negative plateau). For a positive mismatch with respect to the ideal ratio, as in Fig. 14, the duration of the positive jump is larger than that of the negative jump, whereas for a negative mismatch is the other way around. Increasing the mismatch, the duration of the jumps becomes larger while the plateaus shorten and the difference between the positive and negative jumps becomes more apparent. Through this mechanism T/4 have a large degree of flexibility making them more robust to delay-time mismatches than in- and out-of-phase square waves with symmetric duty cycle arising for negative feedback where already for 2% mismatch the transition jumps develop intermediate plateaus [28]. Note that since self- and cross-coupling delays correspond to fiber lengths of several meters mismatches below 1% are easily achievable in practice.

 figure: Fig. 14

Fig. 14 (a)–(d): Robustness of the T/4 dephased fundamental solution changing Tf : 80ns (s0 = 4/3) (a), 81ns (s0 = 1.350) (b), 82ns (s0 = 1.367) (c), and 83ns (s0 = 1.383) (d). P = 1.25, other parameters as in Fig. 11. We also show the frequency of x1 (e) and the phase difference between x2 and x1 (f) as a function of s0. Red dots correspond to numerical simulations of (3) and solid black lines to the theoretical prediction from (9) and (10).

Download Full Size | PDF

 figure: Fig. 15

Fig. 15 As Fig. 14 for the first harmonic with Tf = 80ns (s0 = 4/3) (a), Tf = 80.5ns (s0 = 1.342) (b), Tf = 81ns (s0 = 1.350) (c) and Tf = 81.5ns (s0 = 1.358) (d).

Download Full Size | PDF

Figures 14(e) and Figures 14(f) show the frequency of x1 and the phase difference between x2 and x1 as a function of s0 for the fundamental square wave. The solid line corresponds to the prediction of Eq. (20) whereas red dots have been obtained integrating Eq. (3) numerically and applying the Poincaré section technique. The synchronized fundamental square wave has a practically constant frequency up to 5% mismatch, value at which it becomes unstable and the system jumps to a higher frequency square wave. Within this range the phase difference changes in a practically linear way as predicted theoretically. Higher harmonics are more sensible to mismatches in the delay times. For instance, as shown in Figs. 15(e) and Figs. 15(f), the first harmonic becomes unstable with 2% mismatch. This is in agreement with the linear stability analysis as it can be seen from Fig. 10 where the Hopf bifurcation curves for higher the frequency solutions are sharper. From Fig. 10, we can also argue the reason why the shape of the square waves smoothen: increasing the mismatch while keeping fixed the pump P leads to an increase in Pc along the Hopf branch, leading to a reduction of the distance PPc to the Hopf bifurcation where the solution is sinusoidal. For larger values of P the solutions are less robust to delay-time mismatches and even become chaotic.

8. Final discussion and remarks

We have explored the synchronized square wave solutions generated by two delay-coupled OEOs as function of the offset phases and the ratio between self- and cross-delay times s0. Despite the OEOs being different, synchronized square waves originate from Hopf instabilities of the steady state and multiple harmonics coexist. For both OEO with negative feedback square waves have a symmetric duty cycle and arise supercritically, being synchronized in phase for s0 being an odd/odd rational and out of phase for s0 odd/even rational. For both OEO with positive feedback square waves have an asymmetric duty cycle and are born subcritically. In this case for s0 rational with an odd numerator square waves are in phase while if the numerator is even both in- and out-of-phase square waves coexist. Increasing the pump, subcritical asymmetric square waves are also found beyond the parameter regions where they are born, including regions of negative feedback [29].

We have also shown that for mixed feedback the OEOs show an interesting dynamical regime in which synchronized square waves dephased by a quarter of the period are stable. Furthermore we have shown that this type of synchronization, which does not arise for identical OEOs, exists for a broad range of parameter values and is robust to small mismatches in the delay times.

Altogether this system displays a very large degree of flexibility allowing for instance to select the period of the square waves without changing parameters by choosing a suitable initial condition, to tune the duty cycle while keeping the period constant just changing the offset phase or the pump level and to tune the dephasing between the synchronized square waves by changing the offset phases or the delay times. The methodology and results presented here can be generalized to other systems composed by several oscillators subject to mixed feedback.

Acknowledgments

We acknowledge helpful discussions with Thomas Erneux. We acknowledge financial support of Ministerio de Economía y Competitividad (Spain) and Fondo Europeo de Desarrollo Regional (FEDER) under Projects No. FIS2012-30634 (INTENSE@COSYP) and No. TEC2012-36335 (TRIPHOP), European Social Fund and Govern de les Illes Balears under programs Grups Competitius and Formació de Personal Investigador.

References and links

1. T. Erneux, Applied Delay Differential Equations (Springer, 2000).

2. B. Sartorius, C. Bornholdt, O. Brox, H. Ehrke, D. Hoffmann, R. Ludwig, and M. Mohrle, “All-optical clock recovery module based on self-pulsating DFB laser,” Electron. Lett. 34, 1664–1665 (1998). [CrossRef]  

3. S. Ura, S. Shoda, K. Nishio, and Y. Awatsuji, “In-line rotation sensor based on VCSEL behavior under polarization-rotating optical feedback,” Opt. Express 19, 23683–23688 (2011). [CrossRef]  

4. A. Gavrielides, T. Erneux, D.W. Sukow, G. Burner, T. McLachlan, J. Miller, and J. Amonette, “Square-wave self-modulation in diode lasers with polarization-rotated optical feedback,” Opt. Lett. 31, 2006–2008 (2006). [CrossRef]   [PubMed]  

5. D.W. Sukow, A. Gavrielides, T. Erneux, B. Mooneyham, K. Lee, J. McKay, and J. Davis, “Asymmetric square waves in mutually coupled semiconductor lasers with orthogonal optical injection,” Phys. Rev. E 81, 025206 (2010). [CrossRef]  

6. C. Masoller, D. Sukow, A. Gavrielides, and M. Sciamanna, “Bifurcation to square-wave switching in orthogonally delay-coupled semiconductor lasers: Theory and experiment,” Phys. Rev. A 84, 23838 (2011). [CrossRef]  

7. E.A. Viktorov, A.M. Yacomotti, and P. Mandel, “Semiconductor lasers coupled face-to-face,” J. Opt. B: Quantum Semiclass. Opt. 6, L9–L12 (2004). [CrossRef]  

8. J. Mulet, M. Giudici, J. Javaloyes, and S. Balle, “Square-wave switching by crossed-polarization gain modulation in vertical-cavity semiconductor lasers,” Phys. Rev. A 76, 43801 (2007). [CrossRef]  

9. L. Mashal, G. Van der Sande, L. Gelens, J. Danckaert, and G. Verschaffelt, “Square-wave oscillations in semiconductor ring lasers with delayed optical feedback,” Opt. Express 20, 22503–22516 (2012). [CrossRef]   [PubMed]  

10. X. Zhang, C. Gu, G. Chen, B. Sun, L. Xu, A. Wang, and H. Ming, “Square-wave pulse with ultra-wide tuning range in a passively mode-locked fiber laser,” Opt. Lett. 37, 1334–1336 (2012). [CrossRef]   [PubMed]  

11. K. Ikeda, H. Daido, and O. Akimoto, “Optical turbulence: chaotic behavior of transmitted light from a ring cavity,” Phys. Rev. Lett. 45, 709–712 (1980). [CrossRef]  

12. K. Ikeda and K. Matsumoto, “High-dimensional chaotic behavior in systems with time-delayed feedback,” Physica D 29, 223–235 (1987). [CrossRef]  

13. T. Aida and P. Davis, “Oscillation modes of laser diode pumped hybrid bistable system with large delay and application to dynamical memory,” IEEE J. Quantum Electron. 28, 686–699 (1992). [CrossRef]  

14. K. Ikeda, “Multiple-valued stationary state and its instability of the transmitted light by a ring cavity system,” Opt. Commun. 30, 257–261 (1979). [CrossRef]  

15. A.C. Fowler, Mathematical Models in the Applied Sciences (Cambridge University, 1997).

16. Z. Hong, Z. Feizhou, Y. Jie, and W. Yinghai, “Nonlinear differential delay equations using the Poincaré section technique,” Phys. Rev. E 54, 6925–6928 (1996). [CrossRef]  

17. E. Salik, N. Yu, and L. Maleki, “An ultralow phase noise coupled optoelectronic oscillator,” IEEE Photon. Technol. Lett. 19, 444–446 (2007). [CrossRef]  

18. A. Coillet, R. Henriet, P. Salzenstein, K.P. Huy, L. Larger, and Y.K. Chembo, “Time-domain dynamics and stability analysis of optoelectronic oscillators based on whispering-gallery mode resonators,” IEEE J. Sel. Top. Quantum Electron. 19, 1–12 (2013). [CrossRef]  

19. K. Saleh, R. Henriet, S. Diallo, G. Lin, R. Martinenghi, I.V. Balakireva, P. Salzenstein, A. Coillet, and Y.K. Chembo, “Phase noise performance comparison between optoelectronic oscillators based on optical delay lines and whispering gallery mode resonators,” Opt. Express 22, 32158–32173 (2014). [CrossRef]  

20. S. García and I. Gasulla, “Multi-cavity optoelectronic oscillators using multicore fibers,” Opt. Express 232403–2415 (2015). [CrossRef]   [PubMed]  

21. X.S. Yao and L. Maleki, “Optoelectronic microwave oscillator,” J. Opt. Soc. Am. B 13, 1725–1735 (1996). [CrossRef]  

22. J.P. Goedgebuer, P. Levy, L. Larger, C.-C. Chen, and W.T. Rhodes, “Optical communication with synchronized hyperchaos generated electrooptically,” IEEE J. Quantum Electron. 38, 1178–1183 (2002). [CrossRef]  

23. Y.C. Kouomou, P. Colet, L. Larger, and N. Gastaud, “Chaotic breathers in delayed electro-optical systems,” Phys. Rev. Lett. 95, 203903 (2005). [CrossRef]   [PubMed]  

24. L. Weicker, T. Erneux, O. D’Huys, J. Danckaert, M. Jacquot, Y. Chembo, and L. Larger, “Strongly asymmetric square waves in a time-delayed system,” Phys. Rev. E 86, 055201 (2012). [CrossRef]  

25. L. Weicker, T. Erneux, O. D’Huys, J. Danckaert, M. Jacquot, Y. Chembo, and L. Larger, “Slowfast dynamics of a time-delayed electro-optic oscillator,” Phil. Trans. R. Soc. A 371, 20120459 (2013). [CrossRef]  

26. M. Marconi, J. Javaloyes, S. Barland, M. Giudici, and S. Balle, “Robust square-wave polarization switching in vertical-cavity surface-emitting lasers,” Phys. Rev. A 87, 13827 (2013). [CrossRef]  

27. L. Mei, G. Chen, L. Xu, X. Zhang, C. Gu, B. Sun, and A. Wang, “Width and amplitude tunable square-wave pulse in dual-pump passively mode-locked fiber laser,” Opt. Lett. 39, 3235–3237 (2014). [CrossRef]   [PubMed]  

28. J. Martínez-Llinàs, P. Colet, and T. Erneux, “Tuning the period of square-wave oscillations for delay-coupled optoelectronic systems,” Phys. Rev. E 89, 42908 (2014). [CrossRef]  

29. J. Martínez-Llinàs, P. Colet, and T. Erneux, “Synchronization of tunable asymmetric square-wave pulses in delay-coupled optoelectronic oscillators,” Phys. Rev. E 91, 032911 (2015). [CrossRef]  

30. M. Golubitsky, I. Stewart, P.-L. Buono, and J.J. Collins, “Symmetry in locomotor central pattern generators and animal gaits,” Nature , 401, 693–695 (1999). [CrossRef]   [PubMed]  

31. C. Zhang, R.D. Guy, B. Mulloney, Q. Zhang, and T.J. Lewis, “Neural mechanism of optimal limb coordination in crustacean swimming.,” Proc. Nat. Acad. Sci. 111, 13840–13845 (2014). [CrossRef]   [PubMed]  

32. T. Morbiato, R. Vitaliani, and A. Saetta, “Numerical analysis of a synchronization phenomenon: Pedestrian-structure interaction,” Comput. Struct. 89, 1649–1663 (2011). [CrossRef]  

33. T. Yoshida, L.E. Jones, S.P. Ellner, G.F. Fussmann, and N.G. Hairston Jr., “Rapid evolution drives ecological dynamics in a predator - prey system,” Nature 424, 303–306 (2003). [CrossRef]   [PubMed]  

34. L.E. Jones and S.P. Ellner, “Effects of rapid prey evolution on predator-prey cycles,” J. Math. Biol. 55, 541–573 (2007). [CrossRef]   [PubMed]  

35. T. Hiltunen, L. Jones, S. Ellner, and H.G. Hairston Jr., “Temporal dynamics of a simple community with intraguild predation: an experimental test,” Ecology 94, 773–779 (2013). [CrossRef]  

36. R. Gobbelé, T.D. Waberski, H. Simon, E. Peters, F. Klostermann, G. Curio, and H. Buchner, “Different origins of low- and high-frequency components (600 Hz) of human somatosensory evoked potentials,” Clin. Neurophysiol. 115, 927–937 (2004). [CrossRef]   [PubMed]  

37. S.A. Campbell, R. Edwards, and P. van den Driessche, “Delayed coupling between two neural network loops,” SIAM Appl. J. Math. 65, 316–335 (2004). [CrossRef]  

38. T. Wang, L. Ofman, J.M. Davila, and Y. Su, “Growing transverse oscillations of a multistranded loop observed By Sdo/Aia,” Astrophys. J. Lett. 751, L27 (2012). [CrossRef]  

39. N. Gastaud, S. Poinsot, L. Larger, J.-M. Merolla, M. Hanna, J.P. Goedgebuer, and F. Malassenet, “Electro-optical chaos for multi-10 Gbit/s optical transmissions,” Electron. Lett. 40, 898–899 (2004). [CrossRef]  

40. M. Peil, M. Jacquot, Y.K. Chembo, L. Larger, and T. Erneux, “Routes to chaos and multiple time scale dynamics in broadband bandpass nonlinear delay electro-optic oscillators,” Phys. Rev. E 79, 26208 (2009). [CrossRef]  

41. X.S. Yao and L. Maleki, “Optoelectronic oscillator for photonic systems,” IEEE J. Quantum Electron. 32, 1141–1149 (1996). [CrossRef]  

42. Y.K. Chembo, L. Larger, H. Tavernier, R. Bendoula, E. Rubiola, and P. Colet, “Dynamic instabilities of microwaves generated with optoelectronic oscillators,” Opt. Lett. 32, 2571–2573 (2007). [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 (15)

Fig. 1
Fig. 1 Diagram of the system modeled integrated by two mutually delay-coupled OEOs. Each OEO consists of a Mach-Zehnder interferometer (MZI), a fiber loop with delay time Tf, a photodiode (PD) and a RF amplifier (G) whose output modulates an arm of the MZI. OEOs are fed by a laser diode (LD) whose output is split in two parts by a 50/50 fiber splitter. The OEOs are mutually coupled with cross-feedback delay time Tc.
Fig. 2
Fig. 2 Pc as given by Eqs. (16)(17) for γij = 0.5 and Φ0 = 0. (a) k′ odd and k even, (b) k′ even and k even, (c) k′ odd and k odd, (d) k′ even and k odd. Parameter regions in which Pc is negative or imaginary are plotted in white and grey, respectively.
Fig. 3
Fig. 3 Value of Pc as in Fig. 2 for γ11 = 0.5, γ22 = 0.3, γ12 = 0.2 and γ21 = 0.4.
Fig. 4
Fig. 4 Hopf bifurcations with ω < 10π for Φ1 = 0.2π, Φ2 = 0.3π, Φ0 = 0, γij = 0.5, ε = 4.17 × 10−4 and δ = 1.2 × 10−2. In (b) lines are plotted only in the range where Pc < 2. Red and black lines correspond to in- and out-of-phase oscillations, respectively.
Fig. 5
Fig. 5 Out-of-phase oscillations with symmetric duty cycle for Φ1 = 0.2π, Φ2 = 0.25π, Φ0 = 0, γ11 = 0.5, γ22 = 0.3, γ12 = 0.2, γ21 = 0.4, Tf = 30ns, Tc = 40ns, ε = 6.25 × 10−4, δ = 8 × 10−3 and P = 2.117 (a), P = 2.12 (b), and P = 2.3 (c). Black and red lines correspond to x1 and x2 respectively.
Fig. 6
Fig. 6 Hopf bifurcations for Φ1 = −0.2π and Φ2 = −0.3π. Other parameters as in Fig. 4. The green line corresponds to microsecond oscillations.
Fig. 7
Fig. 7 Fundamental out-of-phase solution for Φ1 = −0.1π, Φ2 = −0.4π, Φ0 = 0, γii = 0.5, Tf = 40ns, Tc = 60ns, ε = 4.17 × 10−4, δ = 1.2 × 10−2, P = 1.702 (a), P = 1.72 (b), and P = 1.8 (c). Pc = 1.7013.
Fig. 8
Fig. 8 Fundamental out-of-phase solution for Φ1 = −0.25π, Φ2 = −0.15π, Tf = 40ns, Tc = 60ns, ε = 4.17×10−4, δ = 1.2×10−2, P = 1.616 (a), P = 1.6271 (just below threshold) (b), P = 1.7 (c), and P = 2 (d). Other parameters as in Fig. 5.
Fig. 9
Fig. 9 Microsecond solution obtained for Φ1 = −0.15π, Φ2 = −0.1π, and P = 1.444768 (a), P = 1.4538 (b), and P = 1.454 (c). Panel (d) shows a zoom of c) close to the maximum. Other parameters as in Fig. 7. The threshold is Pc = 1.4392 and at threshold ω0 = 0.1215. Time traces for x1 and x2 overlap in panels (a)–(c).
Fig. 10
Fig. 10 Hopf lines for Φ1 = −0.25π and Φ2 = 0.15π. Other parameters as in Fig. 4. Yellow and blue lines correspond to bifurcations leading to synchronized solutions dephased +T/4 and −T/4, respectively.
Fig. 11
Fig. 11 Coexistence of T/4 square waves for the parameters of Fig. 10, P = 1.3, and Tc = 60ns. Fundamental, 1 st , 2 nd and 30 th harmonics for Tf = 80ns (s0 = 4/3) are shown in (a)–(d) while (e)–(f) show the corresponding ones for Tf = 40ns (s0 = 2/3). Black and red lines correspond to x1 and x2 respectively.
Fig. 12
Fig. 12 Higher harmonics for the parameters of Fig. 11 but τ = 10ns and s0 = 4/3: (a) 40 th , T = 4/243 (0.99ns), (b) 60 th , T = 4/363 (0.66ns) and (c) 80 th , T = 4/483 (0.50ns).
Fig. 13
Fig. 13 Square-waves dephased T/4 for Tc = 60ns, Tf = 80ns, P = 1.119 (a), P = 1.13 (b), and P = 1.25 (c). Other parameters as in Fig. 11.
Fig. 14
Fig. 14 (a)–(d): Robustness of the T/4 dephased fundamental solution changing Tf : 80ns (s0 = 4/3) (a), 81ns (s0 = 1.350) (b), 82ns (s0 = 1.367) (c), and 83ns (s0 = 1.383) (d). P = 1.25, other parameters as in Fig. 11. We also show the frequency of x1 (e) and the phase difference between x2 and x1 (f) as a function of s0. Red dots correspond to numerical simulations of (3) and solid black lines to the theoretical prediction from (9) and (10).
Fig. 15
Fig. 15 As Fig. 14 for the first harmonic with Tf = 80ns (s0 = 4/3) (a), Tf = 80.5ns (s0 = 1.342) (b), Tf = 81ns (s0 = 1.350) (c) and Tf = 81.5ns (s0 = 1.358) (d).

Equations (23)

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

τ i x ˙ i ( t ) = x i ( t ) θ i 1 y i ( t ) + P { γ i i 2 cos 2 [ x i ( t T i i ) + Φ i ] + γ j i 2 cos 2 [ x j ( t T j i ) + Φ j ] + 2 γ i i γ j i cos [ x j ( t T j i ) + Φ j ] cos [ x i ( t T i i ) + Φ i ] × cos [ x i ( t T i i ) x j ( t T j i ) + Φ i Φ j + ( 1 ) i Φ 0 ] } y ˙ i ( t ) = x i ( t ) ,
x i st = 0 y i st = θ P { γ i i 2 cos 2 Φ i + γ i i 2 cos 2 Φ j + 2 γ i i γ j i cos Φ j cos Φ i cos [ Φ 1 Φ 2 Φ 0 ] } , j i .
ε x i ( s ) = x i ( s ) δ Y i ( s ) + P γ i i 2 { cos 2 [ x i ( s s 0 ) + Φ i ] cos 2 Φ i } + P γ j i 2 { cos 2 [ x j ( s 1 ) + Φ j ] cos 2 Φ j } + 2 P γ i i γ j i { cos [ x i ( s s 0 ) + Φ i ] cos [ x j ( s 1 ) + Φ j ] × cos [ x i ( s s 0 ) + x j ( s 1 ) + Φ j Φ i + ( 1 ) i Φ 0 ] cos Φ i cos Φ j cos ( Φ 1 Φ 2 Φ 0 ) } Y i ( s ) = x i ( s ) ,
s 0 = T f / T c , ε = τ T c 1 , δ = T c θ 1 .
ε U i ( s ) = U i ( s ) δ V i P [ F i U i ( s s 0 ) + K i U j ( s 1 ) ] V i ( s ) = U i ( s ) ,
F i = γ i i 2 sin 2 Φ i + 2 γ i i γ j i cos Φ j sin [ 2 Φ i Φ j + ( 1 ) i Φ 0 ] , K i = γ j i 2 sin 2 Φ j + 2 γ i i γ j i cos Φ i sin [ 2 Φ j Φ i ( 1 ) i Φ 0 ] .
0 = [ 1 + ε ( λ + i ω ) + δ ( λ + i ω ) 1 + P F i e ( λ + i ω ) s 0 ] u i + P K i e ( λ + i ω ) u j ,
0 = [ 1 + ε ( λ + i ω ) + δ ( λ + i ω ) 1 + P F 1 e ( λ + i ω ) s 0 ] × [ 1 + ε ( λ + i ω ) + δ ( λ + i ω ) 1 + P F 2 e ( λ + i ω ) s 0 ] P 2 K 1 K 2 e 2 ( λ + i ω ) .
0 = 1 + P c ( F 1 + F 2 ) [ cos ( ω s 0 ) + ( ε ω δ ω ) 1 sin ( ω s 0 ) ] + P c 2 [ F 1 F 2 cos ( 2 ω s 0 ) K 1 K 2 cos ( 2 ω ) ] ( ε ω δ ω 1 ) 2 ,
0 = P c ( F 1 + F 2 ) [ ( ε ω δ ω 1 ) cos ( ω s 0 ) sin ( ω s 0 ) ] P c 2 [ F 1 F 2 sin ( 2 ω s 0 ) K 1 K 2 sin ( 2 ω ) ] + 2 ( ε ω δ ω 1 ) .
ω s 0 = k π ,
ω = k π / 2.
s 0 = 2 k k .
T = 2 ( 1 s 0 ) 2 k k = 2 s 0 k = 4 k .
1 + P c ( F 1 + F 2 ) ( 1 ) k + P c 2 ( F 1 F 2 K 1 K 2 ) ( 1 ) k ) = 0.
P c = ( 1 ) k F 1 + F 2 ,
P c = ( F 1 + F 2 ) ( 1 ) k ( F 1 F 2 ) 2 + 4 K 1 K 2 ( 1 ) k 2 [ F 1 F 2 K 1 K 2 ( 1 ) k ] .
u 2 = Q u 1 , Q = ( i ) k 1 + P c F 1 ( 1 ) k P c K 1 .
| Q | = [ 1 + ( ε ω δ ω 1 ) 2 + P c 2 F 1 2 + 2 P c F 1 [ cos ( ω s 0 ) ( ε ω δ ω 1 ) sin ( ω s 0 ) ] 1 / 2 P c | K 1 | ,
φ 2 φ 1 = arctan sin ( ω ) + ( ε ω δ ω 1 ) cos ( ω ) + P c F 1 sin [ ω ( 1 s 0 ) ] cos ( ω ) ( ε ω δ ω 1 ) sin ( ω ) + P c F 1 cos [ ω ( 1 s 0 ) ] .
ω k = k π s 0 + α k ,
0 = P c ( F 1 + F 2 ) ( ε α 0 δ α 0 α 0 s 0 ) 2 P c 2 α 0 ( F 1 F 2 s 0 K 1 K 2 ) + 2 ( ε α 0 δ α 0 ) ,
α 0 2 = [ P c ( F 1 + F 2 ) + 2 ] δ P c ( F 1 + F 2 ) ( ε s 0 ) 2 P c 2 ( F 1 F 2 s 0 K 1 K 2 ) + 2 ε .
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.