Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

Time Delay in the Kuramoto Model of Coupled Oscillators

M. K. Stephen Yeung and Steven H. Strogatz


Department of Theoretical and Applied Mechanics,
Kimball Hall, Cornell University, Ithaca, NY 14853-1502
(February 5, 2008)

Nevertheless, we show that perfect synchrony is possi-


We generalize the Kuramoto model of coupled oscillators ble in the Kuramoto model with time delay, if all oscilla-
to allow time-delayed interactions. New phenomena include
tors are identical. In fact, there can be several different
bistability between synchronized and incoherent states, and
unsteady solutions with time-dependent order parameters. synchronized states, and they can co-exist with a stable
We derive exact formulas for the stability boundaries of the incoherent state where the oscillators are completely dis-
arXiv:chao-dyn/9807030v1 22 Jul 1998

incoherent and synchronized states, as a function of the de- organized. These multistabilities are qualitatively new:
lay, in the special case where the oscillators are identical. The they do not occur in the original Kuramoto model.
experimental implications of the model are discussed for pop- We consider a system of phase oscillators with noisy,
ulations of chirping crickets, where the finite speed of sound randomly distributed intrinsic frequencies, and with de-
causes communication delays, and for physical systems such layed mean-field coupling:
as coupled phase-locked loops or lasers.
N
KX
05.45.+b, 02.30.Ks, 87.10.+e θ˙i (t) = ωi + ξi (t) + sin(θj (t − τ ) − θi (t) − α),
N j=1

The Kuramoto model of coupled oscillators is one of (1)


the most celebrated systems in nonlinear dynamics. It
for i = 1, . . . , N. Here θi (t) is the phase of the ith oscilla-
was originally developed as an analytically tractable ver-
tor at time t, and ωi is its intrinsic frequency, randomly
sion of Winfree’s mean-field model for large populations
drawn from a probability density g(ω) with mean ω0 . The
of biological oscillators [1], such as groups of chorusing
white noise ξi (t) represents frequency fluctuations at an
crickets [2], flashing fireflies [3], and cardiac pacemaker
effective temperature D ≥ 0, and is defined by the ensem-
cells [4]. In a beautiful analysis, Kuramoto showed that
ble averages < ξi (t) >= 0, < ξi (s)ξj (t) >= 2Dδij δ(s− t).
the model exhibits a spontaneous transition from inco-
In the global coupling term, K ≥ 0 is the coupling
herence to collective synchronization, as the coupling
strength, τ > 0 is the delay, and α is a phase frustra-
strength is increased past a certain threshold [5]. The
tion parameter. This model reduces to the Kuramoto
model has since been analyzed more deeply and extended
model [5] if τ = 0, α = 0, and D = 0, and to the mean-
in various ways [6–10]. It has also been linked to several
field XY model if τ = 0, α = 0, and the oscillators are
physical problems, including Landau damping in plas-
identical, i.e., g(ω) = δ(ω − ω0 ). For τ = 0, the separate
mas [8], the dynamics of Josephson junction arrays [11],
effects of frustration α and noise D have been studied by
bubbly fluids [12], and coupled Brownian ratchets [13].
Sakaguchi and Kuramoto [6].
Here we explore the effects of time delay on the dynam-
As the one-parameter family of rotating-frame trans-
ics of the Kuramoto model. In the past, delay has often
formations θi (t) → θi (t) − Ωt, ωi → ωi − Ω, α → α + Ωτ
been neglected in models of coupled oscillators. In many
leave Eq. (1) invariant for any Ω, we may assume α = 0
cases this approximation is physically justified, and in
without loss of generality — except if τ = 0, which we for-
all cases it simplifies the mathematics. But recently sev-
bid. (This restriction is merely for convenience. All our
eral authors have begun to investigate oscillator systems
results are well-behaved as τ → 0 and converge to those
where delays are not negligible [14,15], motivated by neu-
obtained by setting τ = 0 from the start.) Moreover,
ral networks where synaptic, dendritic, and propagation
since Eq. (1) is invariant under the reflection ωi → −ωi ,
delays can be significant. Other authors have considered
θi → −θi , α → −α, it suffices to consider ω0 ≥ 0.
delays in systems of limit-cycle oscillators [16], with ap-
It is often helpful to describe the macroscopic state
plications to arrays of lasers and microwave oscillators.
of the system in Pterms of the complex order parameter
Intuitively, the problem is similar to that faced by the N
R(t)eiψ(t) = N1 j=1 eiθj (t) introduced by Kuramoto [5].
fans sitting in an enormous football stadium, all of whom
Here R(t) measures the system’s phase coherence. In par-
(we suppose) are trying to clap in unison. Even if every-
ticular, R = 1 if all the oscillators are in phase, whereas
one were successfully clapping in perfect synchrony, it
R = 0 if the oscillators are scattered around the unit
would not sound that way to the fans themselves, as the
circle with their centroid at the origin.
applause coming from far across the field would be sig-
Our first analytical result concerns the stability of the
nificantly delayed, because of the finite speed of sound.
incoherent state for the infinite-N limit of Eq. (1). We

1
rewrite the model as a Fokker-Planck equation for the Figure 1 shows that this analytical result agrees with
density ρ(θ, ω, t) of oscillators currently at phase θ, with numerical simulations, even for as few as N = 12 oscil-
intrinsic frequency ω. Because the method is standard lators. Although the incoherent state is neutrally stable
[7,8], we omit the details. The only new feature here in the grey region, we observe numerically that R(t) → 0
is that the drift velocity field inherits the time delay exponentially fast, as in Landau damping [8]. In this
in Eq. (1). When the Fokker-Planck equation is lin- sense, incoherence is stable in the grey region.
earized about the incoherent state (the stationary density
ρ ≡ 1/2π), we find [17] that its continuous spectrum is K
{−D − iω | ω ∈ supp(g)}. Hence for D > 0, the continu-
1.5
ous spectrum corresponds to damped modes and there-
fore the stability of the incoherent state is determined
solely by the discrete eigenvalues. But when D = 0, the 1.0
continuous spectrum is pure imaginary and corresponds
to neutrally stable rotating waves in the full system. In
this case, the incoherent state can never be linearly sta- 0.5
ble: it is either unstable or neutral, depending on the
discrete eigenvalues. These eigenvalues λ satisfy [17] τ
Z ∞
−λτ K g(ω) 0 4 8 12 16 20
e dω = 1. (2) FIG. 1. Stability region for the incoherent state, with
2 −∞ λ + D + iω
g(ω) = δ(ω − ω0 ), ω0 = π/2, D = 0, N = 12. Black
This implicit formula for λ is exact but difficult to curves: theoretical boundaries (4) for the infinite-N limit;
analyze for arbitrary g(ω), so we consider the case of Grey area: results from numerical integration using a sixth
order Adams-Bashforth-Moulton scheme, with fixed stepsize
identical oscillators to gain some insight. Even this case
dt = τ /20, and with the corrector formula iterated for con-
turns out to be far from trivial. If g(ω) = δ(ω − ω0 ), Eq. vergence and stability. Initially, all the phases were evenly
(2) can be simplified to the transcendental equation spaced, and the symmetry was broken by adding O(10−10 )
random perturbations. The incoherent state was judged as
(p + ir − z)ez + q = 0, (3) unstable if R(t) > 10−7 at a final time of t = 800τ .

where p = −Dτ ≤ 0, r = −ω0 τ, q = Kτ /2, and z = λτ.


Continuing with the instructive case of noiseless, iden-
Then the stability of the incoherent state depends on
tical oscillators, we now consider the possibility of perfect
whether all roots of Eq. (3) satisfy Re(z) < 0 (in which
synchrony: θi (t) = θ(t) for all i. We restrict our attention
case we will say “all eigenvalues are stable”, for brevity).
to a particular class of such solutions, namely uniform ro-
By the transformations z → z + inπ, q → (−1)n q, r →
tations: θ(t) = Ωt + β. Self-consistency then requires
r + nπ, and z → z ∗ , r → −r, we may assume r ∈ [0, π/2]
in Eq. (3). For r = 0, Hayes proved [18,19] that all Ω = ω0 − K sin(Ωτ ) (5)
eigenvalues are stable if and only if p < 1 and p < −q <
p
p2 + y1 2 , where y1 is the unique zero of p sin y − y cos y for such solutions to exist, and linearization [17] imposes
in (0, π). Using results of Pontryagin as in Ref. [19], we
can show [17,20] that for r ∈ (0, π/2], cos(Ωτ ) > 0 (6)

• if p = 0, then all eigenvalues are stable if and only as the condition for their orbital stability.
if r − π/2 < q < 0. If we graph both sides of Eq. (5) as functions of Ω, we
see that for all sufficiently large K, there exist multiple
• if p p
< 0, then all eigenvaluespare stable if and only stable synchronized states [21], as Eq. (5) has non-unique
if − p2 + (y2 − r)2 < q < p2 + (y1 − r)2 , where solutions satisfying (6). We can also see that stable syn-
y1 and y2 are the unique zeroes of p sin y + (r − chrony is impossible for certain combinations of τ and K.
y) cos y in (0, r) and (π/2, π) respectively. The problem reduces to characterizing the two-parameter
These conditions are exact but still opaque, so we sim- family of lines of negative slope that intersect the sine
plify the model further for illustration. Suppose there is function on its descending limbs. We find that stable
no noise (D = 0). Then we find [17] that the incoherent synchronized states do not exist if and only if
state is neutrally stable precisely when ω0 (4m − 3)π (4m − 1)π
K< and <τ < ,
ω0 (4m − 3)π (4m − 1)π 2(2m − 1) 2ω0 − 2K 2ω0 + 2K
K< and <τ < , (4)
2m − 1 2ω0 − K 2ω0 + K (7)
with m being an arbitrary positive integer. with m being an arbitrary positive integer.

2
These zones of forbidden synchrony are shown in black (a) (b)
1 1
in Fig. 2. For comparison, they are overlaid on top of the R R
earlier grey regions (Fig. 1) where incoherence is stable.
The black regions fit neatly inside the grey; they have the 0.5 0.5
same base and half the height. The exposed parts of the
grey regions correspond to bistability: stable incoherence 0 0
300 350 400 300 350 400
coexists with stable synchrony, and hysteresis can occur. t t
(c) (d)
1 1
K / ω0 R R
1 0.5
0.5

3/4 0 0
300 350 400 0 50 100
t t
FIG. 3. Time series showing nonsteady order parameter
1/2
R(t), with g(ω) = δ(ω − ω0 ), ω0 = π/2, D = 0, τ = 2, N = 24.
(a) K = 1.3 : period-2 oscillation; (b) K = 1.4 : period-4
1/4 oscillation; (c) K = 1.44 : period-8 oscillation; (d) K = 1.475 :
R(t) → 1 after a periodic transient.
ω0τ
0 2π 4π 6π 8π 10π
FIG. 2. Stability regions of the incoherent state ((4)) and The bistability found earlier also has a counterpart in
the synchronized states ((7)) for g(ω) = δ(ω − ω0 ), D = 0. the Lorentzian case (Fig. 4(a)). Partially locked states
White region: one or more stable synchronized states exist (which may not be unique) replace the earlier in-phase
but the incoherent state is unstable; Black region: incoherence states, but otherwise the story is unchanged [23]. Thus,
is stable but synchrony is not; Grey region: one or more stable the case of identical oscillators captures the essential fea-
synchronized states coexist with stable incoherence. tures introduced by delay.
Our final result concerns the bifurcation at K = Kc ,
Numerical simulations reveal windows in the bistable where the incoherent state becomes unstable. We have
regions where R can be time-periodic (Fig. 3). In the ex- adapted the two-timing method of Ref. [9] to handle the
ample shown, period doubling occurs as K increases, but delay-differential equations (1). We find [17] that gener-
seems to be truncated beyond period 16 (not shown); ically, for D ≥ 0 and arbitrary g(ω), a Hopf bifurcation
after that, a synchronized state apparently takes over, occurs at Kc [24], giving rise to a partially locked state, or
suppressing further period doubling (Fig. 3(d)). Such in the density description, a rotating wave with constant
unsteady behavior is a consequence of the delay; in the p
coherence R = O( |K − Kc |). This bifurcation may be
standard Kuramoto model, numerical experiments show subcritical (Fig. 4(a)) or supercritical (Fig. 4(b)).
that R(t) always approaches a constant value if g(ω) is Experimental tests of the model may be possible in
unimodal and symmetric (although this has never been arrays of phase-locked loops [25], relativistic magnetrons
proven). Oscillator configurations with two or more clus- [26], or solid-state lasers [27], as they are all approxi-
ters [22] cause the unsteady behavior seen here. All the mately governed by coupled Adler equations [28] similar
clusters move with the same average velocity, but their in form to Eq. (1). The delay and the coupling strength
separation is periodically modulated. are both natural control parameters, and perhaps one
So far we have concentrated on identical oscillators. could try to map out the stability boundaries, look for
To check how the results would be modified for other fre- hysteresis between incoherence and synchrony, etc. Our
quency distributions, we have considered the Lorentzian model may also help to explain how crickets can synchro-
distribution g(ω) = (γ/π)(γ 2 + (ω − ω0 )2 )−1 . Then by a nize their chirps [2], despite the time delays caused by the
remarkable coincidence [17], Eq. (2) can again be reduced speed of sound. Crickets listen to each other’s chirps and
to Eq. (3), but now with p = −(γ + D)τ, r = −ω0 τ, q = adjust their own timing according to a phase response
Kτ /2, and z = λτ. The critical coupling is given by curve [2]. The propagation delay between two crickets 3
Kc = 2(γ + D) sec(Ωc τ ), (8) meters apart is about 10 msec. This is short compared
to the chirp period (300-500 msec). Our results suggest
where Ωc = ω0 − (γ + D) tan(Ωc τ ). Figure 4 plots the
that delay effects become significant only in the first peak
corresponding region where incoherence is stable. It re-
in Fig. 2, i.e., for delays near half the period of the oscil-
sembles Fig. 1, with a series of evenly spaced peaks (at
lation. Thus the delays that crickets actually encounter
ω0 τ = (2n + 1)π) that decrease in height. The main dif-
in the field are probably negligible as far as synchrony is
ference is that the distributed frequencies produce some
concerned. It would be interesting to try lab experiments
rounding of the boundary, and lift it off the τ -axis so that
on crickets interacting via chirp signals whose delay and
it now has minima 2(γ + D) at ω0 τ = 2nπ.

3
K [11] K. Wiesenfeld, P. Colet, and S. H. Strogatz, Phys. Rev.
4 Lett. 76, 404 (1996); Phys. Rev. E 57, 1563 (1998).
[12] G. Russo and P. Smereka, SIAM J. Appl. Math. 56, 327
1 1 (1996); P. Smereka, preprint (unpublished).
3 [13] R. Häußler, R. Bartussek, and P. Hänggi, in Applied Non-
R R
(a) (b) linear Dynamics and Stochastic Systems near the Mille-
2 0 0 nium, AIP Conf. Proc. Vol.411, edited by J. B. Kadtke
0 2K 4 0 0.5 K 1 and A. Bulsara (Am. Inst. Phys., New York, 1997), pp.
243–248.
1 [14] E. Niebur, H. G. Schuster, and D. M. Kammen, Phys.
Rev. Lett. 67, 2753 (1991).
τ [15] H. G. Schuster and P. Wagner, Prog. Theor. Phys. 81,
0 2 4 6 8 10 939 (1989); B. Dorizzi and B. Grammaticos, Phys. Rev.
FIG. 4. Stability region of the incoherent state for A 44, 6958 (1991); W. Gerstner and J. L. van Hem-
Lorentzian g(ω) with ω0 = 3, γ = 0.1, D = 0, N = 3600. men, Phys. Rev. Lett. 71, 312 (1993); P. C. Bressloff,
Black curves: theoretical boundary Eq. (8); Grey strips: nu- S. Coombes, and B. de Souza, ibid., 79, 2791 (1997);
merical results using the same method as in Fig. 1. Insets: (a) P. C. Bressloff and S. Coombes, ibid., 80, 4815 (1998); S.
Subcritical Hopf bifurcation of the incoherent state at τ = 1. Kim, S. H. Park, and C. S. Ryu, ibid., 79, 2911 (1997); Y.
Arrows indicate a hysteresis loop between stable incoherence Nakamura, F. Tominaga, and T. Munakata, Phys. Rev. E
and a stable partially locked state with R close to 1. (b) Su- 49, 4849 (1994); C. van Vreeswijk, ibid., 54, 5522 (1996);
percritical Hopf bifurcation of the incoherent state at τ = 2. U. Ernst, K. Pawelzik, and T. Geisel, ibid., 57, 2150
A stable partially locked state
√ grows continuously from the (1998); T. B. Luzyanina, Network: Computation in Neu-
incoherent state with R = O( K − Kc ). ral Systems 6, 43 (1995); A. Nischwitz and H. Glünder,
Biol. Cybern. 73, 389 (1995).
[16] J. J. Lynch and R. A. York, IEEE Trans. Circuits Sys.
amplitude can be electronically manipulated. I 42, 413 (1995); S. Wirkus and R. Rand, in ASME De-
Research supported in part by the National Science sign Engineering Technical Conferences (Am. Soc. Mech.
Foundation. We thank Tim Forrest for information Eng., New York, 1997); D. V. R. Reddy, A. Sen, and G. L.
about crickets. Johnston, Phys. Rev. Lett. 80, 5109 (1998).
[17] M. K. S. Yeung and S. H. Strogatz, (to be published).
[18] N. D. Hayes, J. London Math. Soc. 25, 226 (1950).
[19] R. Bellman and K. L. Cooke, Differential-Difference
Equations (Academic Press, New York, 1963).
[20] Physically, p = −Dτ ≤ 0. But see Ref. [17] for the case
p > 0, as well as generalizations of Eq. (3).
[1] A. T. Winfree, J. Theor. Biol. 16, 15 (1967). [21] In Ref. [14] Niebur et al. analyzed the case of large K
[2] T. J. Walker, Science 166, 891 (1969); E. Sismondo, ibid., and small τ for a two-dimensional network with nearest-
249, 55 (1990). neighbor coupling. They found an equation analogous to
[3] J. Buck, Quart. Rev. Biol. 63, 265 (1988). Eq. (5) and that the smallest root corresponded to the
[4] C. S. Peskin, Mathematical Aspects of Heart Physiology most stable solution. This is consistent with (6). However,
(Courant Inst. Math. Sci., New York, 1975). if ω0 τ is not small, the smallest root of Eq. (5) may not
[5] Y. Kuramoto, in International Symposium on Mathemat- be a stable frequency (as dictated by (6)).
ical Problems in Theoretical Physics, Vol. 39 of Lecture [22] D. Golomb, D. Hansel, B. Shraiman, and H. Sompolin-
Notes in Physics, edited by H. Araki (Springer-Verlag, sky, Phys. Rev. A 45, 3516 (1992); K. Okuda, Physica D
Berlin, 1975); Chemical Oscillations, Waves, and Turbu- 63, 424 (1993); K. Kaneko, ibid., 75, 55 (1994).
lence (Springer-Verlag, Berlin, 1984). [23] However, we have not seen unsteady R(t) in simulations
[6] H. Sakaguchi and Y. Kuramoto, Prog. Theor. Phys. 76, for Lorentzian g(ω).
576 (1986); H. Sakaguchi, ibid., 79, 39 (1988). [24] At special parameter values, e.g., the turning points
[7] S. H. Strogatz and R. E. Mirollo, J. Stat. Phys. 63, 613 ω0 τ = nπ of the stability boundary in Fig. 4, Eq. (1)
(1991); J. D. Crawford, J. Stat. Phys. 74, 1047 (1994); is O(2)-symmetric and codimension-2 bifurcations occur.
Phys. Rev. Lett. 74, 4341 (1995). [25] Phase-Locked Loops and Their Applications, edited by
[8] S. H. Strogatz, R. E. Mirollo, and P. C. Matthews, Phys. W. C. Lindsey and M. K. Simon (IEEE Press, New York,
Rev. Lett. 68, 2730 (1992). 1978).
[9] L. L. Bonilla, C. J. Pérez Vicente, and R. Spigler, Physica [26] J. Benford, H. Sze, W. Woo, R. R. Smith, and B. Harte-
D 113, 79 (1998). neck, Phys. Rev. Lett. 62, 969 (1989).
[10] B. Ermentrout, J. Math. Biol. 22, 1 (1985); H. Daido, [27] L. Fabiny, P. Colet, R. Roy, and D. Lenstra, Phys. Rev.
Phys. Rev. Lett. 73, 760 (1994); Physica D 91, 24 (1996); A 47, 4287 (1993).
H.-A. Tanaka, A. J. Lichtenberg, and S. Oishi, Phys. Rev. [28] R. Adler, Proc. IRE 34, 351 (1946), reprinted in Proc.
Lett. 78, 2104 (1997); J. A. Acebrón, L. L. Bonilla, S. D. IEEE, 61, 1380 (1973).
Leo, and R. Spigler, Phys. Rev. E 57, 5287 (1998).

You might also like