Synchronization in Oscillatory Networks: Grigory V. Osipov Jürgen Kurths Changsong Zhou
Synchronization in Oscillatory Networks: Grigory V. Osipov Jürgen Kurths Changsong Zhou
Synchronization in Oscillatory Networks: Grigory V. Osipov Jürgen Kurths Changsong Zhou
I N S YN E R G E T I C S COMPLEXITY
Grigory V. Osipov
Jürgen Kurths
Changsong Zhou
Synchronization
in Oscillatory
Networks
Springer Complexity
Springer Complexity is an interdisciplinary program publishing the best research
and academic-level teaching on both fundamental and applied aspects of complex
systems – cutting across all traditional disciplines of the natural and life sciences,
engineering, economics, medicine, neuroscience, social and computer science.
Complex Systems are systems that comprise many interacting parts with the abil-
ity to generate a new quality of macroscopic collective behavior the manifestations
of which are the spontaneous formation of distinctive temporal, spatial or functional
structures. Models of such systems can be successfully mapped onto quite diverse
“real-life” situations like the climate, the coherent emission of light from lasers,
chemical reaction-diffusion systems, biological cellular networks, the dynamics of
stock markets and of the internet, earthquake statistics and prediction, freeway traf-
fic, the human brain, or the formation of opinions in social systems, to name just
some of the popular applications.
Although their scope and methodologies overlap somewhat, one can distinguish
the following main concepts and tools: self-organization, nonlinear dynamics, syn-
ergetics, turbulence, dynamical systems, catastrophes, instabilities, stochastic pro-
cesses, chaos, graphs and networks, cellular automata, adaptive systems, genetic al-
gorithms and computational intelligence.
The two major book publication platforms of the Springer Complexity program
are the monograph series “Understanding Complex Systems” focusing on the vari-
ous applications of complexity, and the “Springer Series in Synergetics”, which is
devoted to the quantitative theoretical and methodological foundations. In addition
to the books in these two core series, the program also incorporates individual titles
ranging from textbooks to major reference works.
Springer Series in Synergetics
Founding Editor: H. Haken
The Springer Series in Synergetics was founded by Herman Haken in 1977. Since
then, the series has evolved into a substantial reference library for the quantitative,
theoretical and methodological foundations of the science of complex systems.
Through many enduring classic texts, such as Haken’s Synergetics and In-
formation and Self-Organization, Gardiner’s Handbook of Stochastic Methods,
Risken’s The Fokker Planck-Equation or Haake’s Quantum Signatures of Chaos,
the series has made, and continues to make, important contributions to shaping
the foundations of the field.
The series publishes monographs and graduate-level textbooks of broad
and general interest, with a pronounced emphasis on the physico-mathematical
approach.
Hermann Haken
Center of Synergetics, University of Stuttgart, Stuttgart, Germany
Janusz Kacprzyk
System Research, Polish Academy of Sciences, Warsaw, Poland
Scott Kelso
Center for Complex Systems and Brain Sciences, Florida Atlantic University, Boca Raton, USA
Jürgen Kurths
Nonlinear Dynamics Group, University of Potsdam, Potsdam, Germany
Linda Reichl
Center for Complex Quantum Systems, University of Texas, Austin, USA
Peter Schuster
Theoretical Chemistry and Structural Biology, University of Vienna, Vienna, Austria
Frank Schweitzer
System Design, ETH Zurich, Zurich, Switzerland
Didier Sornette
Entrepreneurial Risk, ETH Zurich, Zurich, Switzerland
Grigory V. Osipov Jürgen Kurths
Changsong Zhou
Synchronization
in Oscillatory Networks
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Synchronization Phenomena in Nature, Physics,
and Engineering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Goal of the Book . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Terminological Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Bibliographical Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2 Basic Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1 Harmonic Oscillator: Amplitude, Frequency
and Phase of Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Van der Pol Oscillator: Quasi-Harmonic
and Relaxation Limit Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Rössler Oscillator: From Phase-Coherent
to Funnel Chaotic Attractors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Lorenz Oscillator: “Classic” and Intermittent Chaotic
Attractors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 Phase Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.5.1 First-Order Phase Oscillator (Active Rotator) . . . . . . . 21
2.5.2 Second-Order Phase Oscillator
(Pendulum-Like System) . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5.3 Third-Order Phase Oscillator (Chaotic Rotator) . . . . . 24
2.5.4 Discrete-Time Rotator (Circle Map) . . . . . . . . . . . . . . . . 24
2.6 Discrete Map for Spiking–Bursting Neural Activity . . . . . . . . . . 28
2.7 Excitable Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.7.1 Hodgkin–Huxley Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.7.2 FitzHugh–Nagumo Model . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.7.3 Luo–Rudy Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
X Contents
13.3.3
Noise-Enhanced PS in Arrays of Globally Coupled
Rössler Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
13.3.4 Experimental Observation of Noise-Enhanced PS . . . . 297
13.4 Noise-Enhanced Synchronization-Like Phenomena
in Arrays of Coupled Excitable Cells . . . . . . . . . . . . . . . . . . . . . . 305
13.4.1 Phase Synchrony in Chains of Coupled Noisy
Excitable Neurons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
13.4.2 Noise-Enhanced PS of Coupled Excitable Neurons
by External Forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
13.4.3 Resonant Pattern Formation in 2D Arrays . . . . . . . . . . 313
13.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
Part I
i=1,...,N
j=1,...,M
which the variables of the oscillators are coupled, N and M define the size of
the lattice. The mainly studied boundary conditions are free ends.
In the case of one-dimensional in space structure the network is reduced
to a chain and the model (1.1) can be rewritten as:
ẍ + ω02 x = 0 (2.1)
φ(t) ≡ ω0 t + φ0 (2.3)
ẋ = y,
(2.5)
ẏ = −ω 2 x + μ(1 − x2 )y,
0 0
y
−5
−2
−10
−4 −15
−4 −2 0 2 4 −4 −2 0 2 4
x x
Fig. 2.1. Phase portraits for the van der Pol oscillator (2.5) for μ = 0.12 (a) and
μ = 7 (b)
1
10
0
10
−1
10
−2
10
−3
10
−4 (a)
10 ( b)
−5
10
0.0 0.5 1.0
0.0 0.5 1.0
Ω /2π
Ω / 2π
4.0
0.0
x
(c) (d)
−4.0
100
(e) (f )
50
φ
0 0 50 100
0 50 100
t t
Fig. 2.2. Power spectra (a, b), time series x(t) (c, d), and phases (e, f ) for the
system (2.5) for μ = 0.12 (a, c, e) (quasi-harmonic type) and μ = 7 (b, d, f )
(relaxation type)
0.05
(g) (h) (i)
P(T)
0.00
1 7 1 7 1 7
T T T
Fig. 2.3. Upper panel (a–c): projections of the attractors of the Rössler system
(2.9) onto the plane (x, y); middle panel: (d–f ): projections onto (ẋ, ẏ); lower panel
(g–i): distribution of the return times T . The parameters are ω = 0.98 and a = 0.16
(a, d, g), a = 0.22 (b, e, h) and a = 0.28 (c, f, i)
Hence, we have to use other concepts for an appropriate phase definition [60].
One is based on the general idea of curvature of an arbitrary curve [61]. For
any two-dimensional curve r1 = (u, v) the angle velocity at each point is
ds
ν= /R, (2.10)
dt
where
ds/dt = u̇2 + v̇ 2 (2.11)
is the speed along the curve and
25
15
5
x
−5
−15 (c)
(a) (b)
−25
0 50 100 0 50 100 0 50 100
t t t
10−2
10− 4
0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
Ω/2π Ω/2π Ω/2π
Fig. 2.4. Upper panel (a–c): time series x(t); lower panel: (d–f ) power spectra of
the Rössler system (2.9). The parameters are ω = 0.98 and a = 0.16 (a, d), a = 0.22
(b, e) and a = 0.28 (c, f )
100
10−1
10−2
Dφ
10−3
10−4
10−5
0.15 0.20 0.25 0.30
a
Fig. 2.5. Phase diffusion coefficient Dφ (2.18) for the Rössler system vs. a, ω = 0.98
chaotic time series ẏ(t) (or y(t) in Fig. 2.3a) correspond to a particular
Poincaré section. This means that in this case the phase can be defined
equivalently by examining the maxima or minima of the scalar chaotic
time series without reconstruction of the dynamics in a higher dimensional
phase space and finding a Poincaré section. Note that this approach does
not work always.
Fourth frequency definition. We emphasize that the mean frequency of
chaotic oscillations Ω can be also calculated as
M
Ω = lim 2π (2.20)
t→∞ t
where M is the number of rotations of the phase point around the origin
or the number of crossing of the flow with some Poincaré section during
time t. This method can be directly applied to observed time series, when
one e.g., takes for M the number of maxima of some variable (e.g., y(t))
in (2.9).
Fourth phase definition. The phase in any system can be rather generally
defined by using the Hilbert transform. This approach is based on the
analytic signal concept [71] and was introduced by Gabor [70]. Given a
signal s(t), the analytic signal ζ(t) is a complex function of time defined as
(here the integral is taken in the sense of the Cauchy principal value).
The instantaneous amplitude A(t) and the instantaneous phase φ(t) of
the signal s(t) are thus uniquely defined from (2.21).
To conclude this section we want to note that for some typical classes of oscil-
lators, e.g. phase-coherent oscillators, the phases and the frequencies defined
in different ways are practically equivalent, and they give the same results in
the study of phase synchronization. However, for some other classes of sys-
tems, e.g., oscillators with funnel attractors, only one or two of the phase and
frequency definitions are applicable.
50 50
(a)
25 25
z
0 0
−20 0 20 −5 15 35
x u
Projections of the “classical” Lorenz attractor on the planes (x, z) (a) and
Fig. 2.6.
(u = x2 + y 2 , z) (b). σ = 10, b = 8/3, and r = 28
ẋ = −σ(x − y),
ẏ = (r − z)x − y, (2.23)
ż = −bz + xy,
0
x
−60
100
0
y
−100
250
z
50
0 10 20 30 40 50 60 70 80 90 100
time
Fig. 2.7. Intermittent chaotic oscillations in the Lorenz oscillator for σ = 10, b =
8/3, and r = 166.1
laminar stage begins. The average length of the laminar stage (ALLS) for
any intermittent type-I chaotic oscillator is calculated as [72, 73]:
1
τ0 ∝ √ , (2.26)
r − rcr
where r is a bifurcation parameter and rcr is the critical value when chaos
sets in1 . We use now the onset of each laminar epoch tn as a marker event
in order to introduce a phase of these intermittent oscillations. Then the
interval [tn , tn+1 ], which covers the whole cycle of one laminar and one
turbulent stage, is attributed to a 2π phase increase and in between we
use a linear interpolation (see also 2.19):
t − tn
φ(t) = 2π + 2πn, tn ≤ t < tn+1 . (2.27)
tn+1 − tn
1
Note, that often because of τ /T 1 the time of the full cycle Tc = τ + T , i.e., the
time between the beginnings of two sequential laminar stages, practically equals
to τ . Therefore, the coincidence of the averaged τ leads to the coincidence of the
averaged Tc .
2.5 Phase Oscillators 21
0.3
(a)
0.2
0.1
x(n)
0
−0.1
−0.2
0.3
0.2
(b)
0.1
x(n)
0
−0.1
−0.2
Here g regulates the coherence properties of the chaotic attractor: for g < 5
the laminar stage duration is distributed in a rather narrow band, i.e. the
chaotic behavior is highly coherent, but for g > 5 this distribution is rather
broad (Fig. 2.8).
180
160
140
ω =2
120
100 ω =1.5
φ
80
60
ω =1.1
40
20 ω =1.01
0
0 10 20 30 40 50 60 70 80 90 100
time
Fig. 2.9. Phase variable evolution in (2.30) for different ω
1.2
D3
1.0
0.8 D2
ω
0.6
ωL
D1
0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
λ
Fig. 2.10. Parameter plane (ω, λ) showing different types of behavior of (2.31)
24 2 Basic Models
where φk is the phase variable at discrete times k = 1, 2, ...; ω ∈ [0; 2π] can
be interpreted as frequency; F (φ) is a piecewise linear 2π-periodic function of
the form:
F (φ) = cφ/π (2.34)
defined in the interval [−π, π], and c ∈ [−π, π] is the control parameter. One
can use any 2π-periodic function F (φ), e.g., F (φ) = sin φ; but our choice (2.34)
of a piecewise linear function F (φ) is motivated not only by the simplicity of
consideration (see [4, 75]) but also by the requirement that there exists chaos
for c < 0 and there are no stable periodic orbits for any c < 0.
First, we shortly describe basic properties of this circle map. It has for
ω < |c| a unique fixed point φ̄ = ωπ/c which is stable if φ̄ ∈ [0; π] and
unstable if φ̄ ∈ [−π; 0].
The dynamics of a CM can be mainly described by the rotation number
ρ, which is defined as the average growth rate of the phase (compare with
(2.16)):
1 φM − φ1
ρ= lim , (2.35)
2π M →∞ M
where M is the number of iterations. (2.35) is valid for regular (c ≥ 0) and
chaotic c < 0 dynamics and defines characteristic time scale of rotations. The
regions of some rational ρ on the plain (c, ω) are marked in Fig. 2.12 with
different levels of gray color.
The parameter c controls the coherence properties of the motions. As a
measure of the degree of coherence analogous to (2.18), we use the variance
D of the phase dynamics which is defined for large k as:
γ = Nd /Ng . (2.37)
Then, we easily see that for c > −ω, γ = 0, but otherwise γ = 0 (Fig. 2.15).
System (2.33) is one of the basic models in nonlinear dynamics, and it
has been studied in many mathematical (cf., [76]), physical (cf., [77–79]) and
engineering issues (in particular, in the theory the digital phase-locked loops
[57, 74, 81]).
26 2 Basic Models
6.283
1
6/7
5/6
4/5
3/4
2/3
3/5
1/2
2/5
1/3
1/4
1/5
1/6
1/7
0 0
c
Fig. 2.12. Distribution of rotation numbers of the circle map (2.33) on the plain
(c, ω). Several regions where the rotation numbers are rational (ρ = p/q) are pre-
sented. From bottom to top different gray level regions are ordered as shown on
the right side. Region in which ρ = 0 is white. Region in which ρ = 1 is black.
Between these regions there exist (but not presented) relatively small regions with
other rational rotation numbers
5.0
0.6
1.0
1.2 1.4
4.0 1.8
2.2
2.6
0.8
3.0
D
0.4
−1.9 −1.7 −1.5 −1.3
2.0
1.0
0.0
−3.0 −2.0 −1.0 0.0 1.0 2.0 3.0
c
Fig. 2.13. The variance D of φk (2.36) vs. c for the map (2.33) at different values
of the frequency parameter ω
2.5 Phase Oscillators 27
180.0
c = 0.0
−0.5
−3.0
0.5
120.0 −2.5
phase
−1.0
−1.5
60.0 −2.0
0.0
0 50 100 150 200 250 300
k
Fig. 2.14. Phase variable evolution at ω = 0.6 for different c for the map (2.33)
1.5
0.6
1.0
1.4
1.8
2.2
2.6
1.0
γ
0.5
0.0
−3.0 −2.5 −2.0 −1.5 −1.0 −0.5 0.0
c
Fig. 2.15. Ratio γ (2.37) of the duration of phase decreasing intervals to the dura-
tion of phase increasing intervals in dependence on c for different values ω in the
map (2.33)
28 2 Basic Models
where x(n) and y(n) are, respectively, the fast and slow dynamical variables.
The slow evolution of y(n) is due to small values of the positive parameters β
and σ (each one of the order 10−3 ). The parameter α controls the dynamics
of the fast variable x(n). Typical regimes of temporal behavior of the map
(2.38) are shown in Fig. 2.16.
The bursting activity is characterized by a distinct timescale which allows
to introduce the phase and the frequency of bursting in each oscillator. The
phase of bursting oscillations φ(n) increases linearly between the moments nk
at which the kth burst starts and gains a 2π growth over each time interval
nk+1 − nk
n − nk
φ(n) = 2πk + 2π (2.39)
nk+1 − nk
(see 2.19 and (2.27)). The mean frequency of the burst dynamics is an average
speed of the phase increase (see 2.17)
φ(n) − φ(0)
Ω = lim . (2.40)
n→∞ n
2.7 Excitable Systems 29
(a)
1.5
0.5
x(n)
−0.5
−1.5
−2.5
(b)
1.5
0.5
x(n)
−0.5
−1.5
−2.5
3000 4000 5000
n
Fig. 2.16. Typical waveforms of spiking–bursting behavior generated by the map
(2.38) for: (a) α = 4.1 and (b) α = 4.4, and β = σ = 0.001
20
−20
V (mv)
−40
−60
−80
0.0 2.0 4.0 6.0 8.0 10.0
t
Fig. 2.17. Typical time series in HH model (2.41)
facial cold receptors and hypothalamic neurons of the rat, etc. The equations
of this specific HH model read:
dV
CM = −Il − Id − Ir − Isd − Isr + Dξ(t)
dt
ar φ(T )(ar∞ − ar )
= ,
dt τr
asd φ(T )(asd∞ − asd )
= ,
dt τsd
asr φ(T )(−ηIsd − θasr )
= , (2.41)
dt τsr
with Id = ρ(T )gd ad∞ (V −Vd ), and Ik = ρ(T )gk ak (V −Vk ), (k = r, sd), where
(T −T0 )/10 (T −T0 )/10
ak∞ = [1 + exp(−sk (V −V0k ))]−1 and ρ(T ) = A1 , φ(T ) = A2 .
Here V is the membrane potential, and Il is the leakage current. ak is the
activation variable, and ρ(T ) and φ(T ) are temperature-dependent scaling
factors. A more detailed description of the model, its parameters, such as
CM , gk , τk . . ., and comparison with the experimentally observed temperature
dependence of spike train patterns can be found in [464]. It has been shown
that this system exhibits a homoclinic bifurcation, where the interspike inter-
val becomes very long, when the control parameter T , the temperature, is
varied [465]. While the classical HH model can only display periodic spiking
behavior, this HH model of thermally sensitive neurons can display chaotic
spiking–bursting behavior as illustrated in Fig. 2.17.
x3
ẋ = F (x, y) = x +− y,
3 (2.42)
ẏ = εG(x, y) = ε(ax + by − c),
2.7 Excitable Systems 31
Fig. 2.18. The original Nagumo circuit is held as a trust in the laboratory of Dr.
K. Aihara in University of Tokyo. Photo provided by H. Suetani and K. Aihara with
permissions.
1.5
1
E D
0.5
h_(y)
ho(y) h+(y)
0
Y
−0.5
C
A
B
−1
G(x,y)=0 F(x,y)=0
−1.5
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
X
Fig. 2.19. Typical phase diagram for the excitable FitzHugh–Nagumo system,
showing the null-clines F (x, y) = 0 (dashed line) and G(x, y) = 0 (solid line), and
stochastic trajectories produced by the system (2.42) in the presence of an additive
noise term in the first equation
3.0
2.0
1.0
0.0
x
−1.0
−2.0
−3.0
0.0 10.0 20.0 30.0 40.0 50.0
t
Fig. 2.20. Typical time series in FHN model (2.42) in the presence of an additive
noise term in the first equation
In the phase plane (x, y) we can hence distinguish two main types of
motion:
(1) Activation, i.e., motion between the null-cline G(x, y) = 0 in the vicinity
of the steady state and the line x = 0 (approximately between the points
A and B)
2.7 Excitable Systems 33
(2) Excursion, i.e., motion after the first passage of the line x = 0 along the
branches h+ (y) and h− (y) till the first passage of the null-cline G(x, y) = 0
(approximately from point B through points C, D, E, to point A).
as specified in the Luo–Rudy phase I model [94]. For simplicity, we will call
system (2.43) Luo–Rudy model. The form of the sodium current Ina , the slow
inward calcium current Isi , and the potassium current Ik , is given by:
where
g∞ = αgi /[αgi + βgi ] (2.47)
is the steady-state value, τgi = 1/[αgi + βgi ] is the time constant, and the
α’s and β’s are functions of the membrane voltage. The form of the time-
independent potassium current Ik1 , the plateau potassium current Ikp , and
the background current Ib , is given by
Ii = Ai (V − Ei ), (2.48)
where Ei is the reversal potential of the ion and Ai is a scaling factor. Note that
Ai is a function of voltage for Ik1 and Ikp . The conductances (Gi (mS/cm2 ))
for each current are the following: Gna = 23, Gsi = 0.07, Gk = 0.705, Gk1 =
0.6047, Gkp = 0.0183, Gb = 0.03921. The reversal potentials (Ei (mV)) are:
Ena = 54.44, Ek = −77, Ek1 = −87.23, Ekp = −87.23, Eb = −59.87. The
reversal potential of Isi depends upon the internal calcium concentration,
which varies with time. More details can be found in [94].
34 2 Basic Models
40
20
−20
V
−40
−60
−80
−100
0 20 40 60 80 100
time
Fig. 2.21. Typical response of the Luo–Rudy model (2.43)–(2.48) on a super-
threshold force
It is clear that a single cell without any perturbation, has a stable steady
state. If some subthreshold force is applied (in (2.43) Istimulus = 0), the system
becomes excited (depolarization stage). After the excitation the system comes
back to the steady state (repolarization stage). The response of the single cell
has a pulse-like form as illustrated in Fig. 2.21.
The Luo–Rudy phase 1 model has been constructed based on experimental
data. Due to its high accuracy in reconstructing the membrane potential and
the ion currents, it is widely used in numerical experiments. Two- and three-
dimensional simulations with the Luo–Rudy model allows to reproduce many
known features, such as single spiral wave or spiral wave chaos, observed in
heart tissue (see special issue [393]).
3
Synchronization Due to External
Periodic Forcing
the paradigmatic examples of the Rössler (Sect. 3.2) and Lorenz (Sect. 3.3)
oscillators, respectively. A detailed description of the phase synchronization
mechanism is presented in Sect. 3.4. Section 3.5 is devoted to the transi-
tion to phase synchronization in chaotic intermittent oscillators. Synchronous
response of excitable systems is discussed in Sect. 3.6. In Sect. 3.7 we conclude
this chapter.
SELF−
EXTERNAL PERIODIC FORCE SUSTAINED
OSCILLATOR
Amplitude A Mean
Frequency ω frequencyω0
Fig. 3.1. Oscillator under external periodic force. In synchronous regime the fre-
quency Ω of driven oscillator is equal to the frequency of external force ω
3.1 Synchronization of Limit-Cycle Oscillator by External Force 37
Introducing the difference between the phase of the oscillations φ and the
phase ωt of the external force:
θ = φ − ωt (3.4)
θ̇ = −Δ + εq(θ), (3.5)
θ̇ = −Δ + ε sin θ. (3.6)
This first-order phase equation has been described in Sect. 2.5.1. Its stable
steady-state solution
Δ
θ̄ = arcsin (3.7)
ε
exists if
|Δ| < |ε| (3.8)
and results in a constant phase shift (strict phase locking) between the phases
of the driving and the driven (the whole) systems. Therefore, external syn-
chronization takes place if condition (3.8), which defines the synchronization
area (the band of synchronization), is fulfilled and is defined through the ratio
of the frequency mismatch Δ and the amplitude A of the external force. The
transition from synchronization regime to beating regime, which takes place
for Δ = A, occurs through the collision and disappearance of the stable and
unstable steady states.
Now we treat an arbitrary (not small like in Sect. 3.1.1) external force applied
to a limit-cycle oscillator, namely to a van der Pol oscillator. From the various
studied cases we present results of an external synchronization on a weakly
nonlinear van der Pol oscillator (2.4). The model system is then:
ẋ = y,
(3.9)
ẏ = −ω02 x + μ[(1 − x2 )y + 2A cos ωt],
1.2
D D
1.0
B
0.8
0.6
A
0.4 C C
O
0.2
0.0
−2.0 −1.5 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0
Δ
still exists but only in the average, i.e., the difference between the phase
of the oscillations and the phase of the external force is not constant but
bounded. By the transition D → C, a point on the limit cycle becomes
rotating around the origin, i.e., a beating regime occurs. This transition
can be better understood by a bifurcation analysis of system (3.10) rewrit-
ten in the real coordinates R and θ (a = R(t)exp[iθ(t)]). In the cylindrical
phase space (R, θ) by the transition D → C the oscillatory limit cycle
disappears but a rotatory limit cycle appears – the cycle which envelops
the cylinder.
20 0.5
10 0.4
0.3
0
A
x
0.2
−10 0.1
(a) (b)
−20 0.0
0 50 100 0.95 0.97 0.99 1.01
time ω
Fig. 3.3. (a) Synchronization of periodic oscillation (solid line) to a weak periodic
driving signal (dotted line). (b) Arnold tongue of the 1 : 1 synchronization
40 3 Synchronization Due to External Periodic Forcing
(a) (b)
10
y 0
−10
−20
−20 −10 0 10 20 −20 −10 0 10 20
x x
0.20
A
0.10
0.00
0.98 1.00 1.02
ω
Rössler oscillator is phase coherent (see Sect. 2.3, Figs. 2.3 and 2.4). Here the
stroboscopic technique is very useful. Its idea is the following: We observe the
variable(s) of the driven system only at the moments tk = kT , where T is
the period of the external force and k = 1, 2, . . .. As shown in Fig. 3.4, when
the system is phase locked to the driving signal, the stroboscope of the sys-
tem state (x, y) at each period of the driving signal is restricted to an arc
area of the chaotic attractor, while it is distributed relatively uniformly over
the whole attractor when the system is out of the phase locking region. The
whole synchronization region, shown in Fig. 3.5, is very similar to the Arnold
tongue of the periodic oscillators in Fig. 3.3b. These properties were firstly
reported in [109] and studied more intensively in [110]. In the marked region
the mean frequency of the chaotic oscillations is equal to the frequency of the
external force, and the difference between the phases is bounded for all time.
This example shows that by using of a periodic external force, we are able to
control the spectral properties of complex signals.
Intuitively, we can expect this similarity because we have noted that
the phases of phase-coherent chaotic oscillations increase almost linearly
3.2 Phase Synchronization of a Chaotic Rössler Oscillator 41
ẋ = 10(y − x),
ẏ = rx − y − xz, (3.12)
ż = xy − 2.667z + A cos ωt.
The unforced Lorenz system exhibits rich bifurcations when the parameter
r is varied [61]. For r = 210, chaotic oscillations result from a periodic-doubling
scenario, and, as mentioned above, the system can then be synchronized per-
fectly by a periodic driving signal with a frequency close to the average fre-
quency Ω = 24.92 of the chaotic oscillations, as it can be seen by the plateau
of the vanishing frequency difference Ω − ω = 0 in Fig. 3.6. The situation
becomes more complicated for r = 28 where a certain plateau of Ω − ω = 0
appears; however, this plateau is neither perfectly horizontal nor lies at zero.
As a result, PS in this case is not perfect, as shown by 2π or 4π phase slips
in Fig. 3.7.
The reason for this imperfect phase synchronization lies in the broad dis-
tribution of the timescales of the UPOs at r = 28. For this parameter, there is
a saddle point (0, 0, 0) embedded (situated very close) in the chaotic attractor.
The trajectory slows down considerably when passing near the saddle point,
ω (r = 210)
0.1
0.05
Ω−ω
−0.05 ω (r = 28)
Fig. 3.6. Perfect and imperfect phase synchronization for periodically forced Lorenz
systems (3.12); solid line: r = 28, A = 6; dotted line: r = 210, A = 3
3.3 Imperfect Phase Synchronization 43
16π
5895 5900
12π
2πN(t) - ωt
8π
4π
0
0 2000 4000 t 6000 8000
Fig. 3.7. Phase slips of imperfect phase synchronization in the periodically driven
Lorenz system at r = 28 (3.12)
Fig. 3.8. Individual frequencies of unstable periodic orbits embedded into the
Lorenz attractor at r = 28 (3.12); dashed line: mean frequency of the autonomous
chaotic motion
while the oscillation is much quicker when the trajectory rotates around one of
the two unstable foci [61]. Thus, the timescales have a relatively large variation
around the average value, as can be seen in the distribution of the frequency
of different UPO (Fig. 3.8). Due to this large variation of the frequencies,
the phase locking regions of the UPOs do not all overlap to produce a full
PS region of the chaotic attractor. For a given frequency and amplitude of
the external driving signal, there exist certain unstable orbits which are not
locked and imperfect phase synchronization characterizes this complicated
locking behavior.
44 3 Synchronization Due to External Periodic Forcing
6
A
0
8.25 ω 8.5
Fig. 3.9. Overlapping of phase locking regions of different locking ratios in the
periodically driven Lorenz system at r = 28 (3.12). Solid line: l = 7 (1 : 1), dashed
line: l = 15 (14 : 15); dotted line: l = 20 (18 : 20)
3.4 Transition to the Regime of Chaotic Phase Synchronization 45
which is essentially the circle map coupled to a chaotic map f (see Sect. 2.5.4).
In this presentation, Ω denotes the difference between the natural frequency
of the chaotic oscillator and the frequency of the driving force, is the cou-
pling strength which is proportional to the amplitude of the driving force,
and g(x(n)) corresponds to the nonuniformity of the phase rotations in the
chaotic oscillators as a result of chaotic fluctuations of the amplitude x. The
average growth rate of the phase, i.e., ΔΩ = limn→∞ [φ(n) − φ(0)]/n, corre-
sponds to the phase rotation number, and ΔΩ = 0 indicates synchronization
of the chaotic oscillator to the external force. Without loss of generality, the
periodically forced chaotic tent map f (x, φ) = 1 − 1.9|x| + 0.05 sin[2πφ(n)]
and g(x) = 0.05x were studied in [117].
The analysis of (3.14) is based on the presentation of a chaotic attractor
through its UPOs embedded in it [79]. A periodic orbit of period N has its
real period T ≈ T0 N , where T0 is the average return time of the period
“1” periodic orbit. For different periodic orbits, T0 shows fluctuations around
the average return time of the chaotic oscillations, as is modeled in the map
system by the term g(x). Due to these fluctuations, each periodic orbit has its
individual phase locking region under the periodic external forcing (Fig. 3.10).
In this illustrative mapping model, the region of complete PS is given by the
overlapping region of the synchronization regions (in the forms of Arnold
tongues) of all the UPOs.
Now we consider, how the transition from phase synchronization regime
looks from the viewpoint of the bifurcation theory. In the PS region, and for
a particular UPO, the uncoupled circle map (in (3.14 g(x(n)) = 0) has a
stable fixed point φs and an unstable fixed point φu , and accordingly, each of
the unstable periodic orbit is associated with an attractor and a repeller in
the direction of φ. In the generalized phase space (x, φ), the attractor (x, φs )
and the repeller (x, φu ) are well separated, as shown in Fig. 3.11a for the
mapping system (3.14). In this generalized phase space with the unwrapped
phase variable (the phase variable is viewed on the real line rather than on
the circle, and phase points separated by 2π are not regarded as the same).
The repellers are periodic orbits on the basin boundary of the attractors [118].
46 3 Synchronization Due to External Periodic Forcing
Fig. 3.10. Phase locking regions for periodic orbits with period 1–5 in system
(3.14). The region of full phase synchronization regime, where all the phase locking
regions overlap, is delineated with black
Fig. 3.11. Stable (in the φ direction) (pluses) and unstable (filled circles) peri-
odic orbits with periods 1–8 forming the skeletons of the attractor and repeller,
respectively (3.14). (a) Inside the full phase locking region, where the attractor and
the repeller are distinct. (b) Just beyond the border of the phase locking region at
which the attractor and repeller collide, the chaotic attractor as a whole is no longer
attractive in the φ direction
When the parameter moves close and crosses the phase locking boundary, the
attractor and the repeller of a few UPOs approach to each other, coalesce and
annihilate as a result of a saddle-node bifurcation, as illustrated in Fig. 3.11b.
Hence, these UPOs are not locked by the external force and phase slips may
3.5 External Phase Synchronization of Chaotic Intermittent Oscillators 47
occur. The dynamics in the weakly unstable direction φ is the same as in the
usual saddle-node bifurcation.
We emphasize that this desynchronization bears much similarity to the
riddling bifurcation in coupled identical systems where a certain UPO becomes
unstable in the transverse direction [120].
The scaling properties obtained in the transitions to (from) the PS regions
will be discussed in Sect. 4.2.1.
Let us start with a well-studied 1D map that exhibits the type-I intermittency
route to chaos (see Sect. 2.4) under external periodic forcing:
where A and ω are the amplitude and the frequency of the external force,
respectively, and f (x) consists of the standard quadratic part and a somewhat
arbitrary chosen return part that acts as a turbulent stage and ensures chaos
ε + x + x2 , if x ≤ 0.2,
f (x) = (3.16)
g(x − 0.2) − ε − 0.24, if x > 0.2.
48 3 Synchronization Due to External Periodic Forcing
First we will focus on the case of a coherent chaotic attractor and let g = 2.
We remind, that without external force the map (3.15) demonstrates a type-I
intermittent behavior for ε > 0, i.e., εcr = 0.
As mentioned before, ALLS τ can be viewed as the characteristic time
scale in systems with intermittent behavior. In order to determine ALLS,
we rewrite the map in the form of a first-order time-continuous differential
equation, which is obtained from (3.15) and (3.16) [73, 130]:
ẋ = ε + x2 + A cos ωt (3.17)
Using the following change of the variable x = −u̇/u, we obtain the Mathieu
equation:
ü + (ε + A cos ωt)u = 0. (3.18)
We will study in (3.18) the well-known cases of parametric resonance.
Parametric resonance of the kth order can be achieved when the following
relation is maintained [132]:
√ k
ε ≈ ω, k ∈ N (3.19)
2
Then the solution is characterized by harmonic oscillations with the frequency
k
2 ω and the exponentially growing amplitude
k
u = a cos ( ωt + φ)epk t , (3.20)
2
where a and φ are some constants, and pk depends upon the number of the
zone of the parametric resonance and the parameters of the systems. After
transformation to the original variable, the parametric instability vanishes
and one gets x = k2 ω tan ( k2 ωt + φ) − pk , which yields for the ALLS in (3.15)
and (3.16) in the synchronous regime:
2
τs ∝ (3.21)
kω
and finally: √
2 ε τ0
τs = , (3.22)
kω
where τ0 corresponds to the case of the autonomous map. So, inside a zone of
parametric resonance that takes place in (3.18), the exponential growth does
not affect the solution of (3.17), which is our concern, and all that matters is
the frequency of the solution. On the opposite, being outside of a resonance
zone, one gets a two-frequency solution in (3.18) and no synchronization exists
k
in the original system (3.17): u = aei 2 ωt eiΩk t + c.c. For the first Mathieu zone
(k = 1) the boundaries are given by
ω
A = 4ε| √ − 1|. (3.23)
2 ε
3.5 External Phase Synchronization of Chaotic Intermittent Oscillators 49
(a) (b)
2000 90
ε=2.64*10−6
3.77*10−6
4.9*10−6 70 A=1.9*10−6
1800
50
φ2−φ1
<τ>
30
A=2.1*10−6
1600
10
A=2.3*10−6
1400 −10
0 1*10−6 2*10−6 3*10−6 0 150000 300000
A time
Fig. 3.12. (a) Locking of the ALLS by external periodic driving in (3.15) and (3.16).
(b) Phase difference evolution in nonsynchronous (A = 1.9 × 10−6 , 2.1 × 10−6 ) and
synchronous (A = 2.3 × 10−6 ) regimes in (3.15) and (3.16) for ε = 2.64 × 10−6
1
Note
√ that from the last equation we have the ALLS for autonomous map < τ0 >≈
π/ ε.
50 3 Synchronization Due to External Periodic Forcing
(a) (b)
4*10−5 6*10−6
3*10−5 S3
4*10−6 S1+ S1−
2*10−5
S2
ε
ε
2*10−6
1*10−5
S1
0
Ioff 0
Ioff
−1*10−5
0 1*10−5 2*10−5 2*10−6 0 2*10−6 4*10−6
A A
Fig. 3.13. (a) Three first zones of synchronization (Sk , k = 1, 2, 3) and the region
where intermittency is absent (Iof f ) in (3.15) and (3.16). (b) Regions of synchroniza-
tion (the first zone) with a positive (S1+ ) and a negative (S1− ) Lyapunov exponent
(LE). The theoretical border of the first synchronization zone (3.23) is the curve
marked by “o”
is not synchronized. If one measures ALLS by taking into account only the
long timescale and neglecting the fast passages, synchronization by external
driving is clearly observed. It allows to claim the existence of imperfect phase
synchronization, if the driven system behaves nonphase-coherent [112].
Now we present the results of numerical simulations of the Lorenz system (that
also exhibits type-I intermittency for r ≈ 166.06 [71]) under multiplicative
external driving which may also be regarded as a modulation of the bifurcation
parameter r [133]:
⎧
⎨ ẋ = σ(y − x),
ẏ = −y − xz + (r + A) cos ωt, (3.25)
⎩
ż = −bz + xy,
0.025
r−rc = −0.0005
0.020 0.0005
0.0015
0.015
1/<τ>
0.010
0.005
0.000
0.000 0.005 0.010 0.015 0.020
A
Fig. 3.14. Synchronization plateaus for the Lorenz system under external driving.
Here rc = 166.06149, τ1 = 0 corresponds to nonintermittent motion
52 3 Synchronization Due to External Periodic Forcing
(a)
50
Istimulus
30
10
−10
(b)
50
voltage
−50
−100
(c)
50
voltage
−50
−100
3000 3500 4000 4500 5000
time
Fig. 3.15. Synchronous response of the Luo–Rudy system on the external peri-
odic stimulation. (a) The train of external pulses. (b) Voltage evolution for 2:1
synchronous excitation. (c) Voltage evolution for 1:1 synchronous excitation. The
parameters are amplitude of the external stimuli =50, pulse duration = 10 ms. The
initial conditions are different
3.7 Conclusions 53
0.07
0.06
1:1
0.05
Ω
0.04
2:1
0.03
0.02
0.045 0.050 0.055 0.060 0.065 0.070 0.075 0.080
ω
Fig. 3.16. Hysteresis by the synchronous response of the Luo–Rudy system on the
external periodic stimulation. The response frequency Ω vs. the input frequency ω
for parameters as in Fig. 3.15 is plotted
3.7 Conclusions
θ̇ = Δ − 2εq(θ), (4.3)
where Δ = ω2 − ω1 and
Δ
q(θ) = (4.5)
2ε
correspond to a synchronous behavior in (4.2) and hence also in the original
system (4.1). Note that here we have not made the restriction to weakly non-
linear oscillators. Therefore, model (4.3) works for relaxation oscillators as
1
Here and in other chapter we use letter ε if the coupling between elements is
small.
4.1 Synchronization of Regular Systems 57
well [153]. In the latter case the function q(θ) may have a discontinuity at the
origin that leads to the rapid appearance of an in-phase (θ ≈ 0) synchroniza-
tion regime.
The simplest coupling function is the sine, i.e., q1 (θ) = −q2 (θ) = sin θ.
Then the phase model (4.3) is:
θ̇ = Δ − 2ε sin θ, (4.6)
i.e., the first-order Adler equation (for a detailed description of this equation
see Sect. 2.5.1)
Its steady state
Δ
θ̄ = arcsin (4.7)
2ε
exists if
|Δ| < |2ε| (4.8)
and is stable. It defines a constant phase shift between the phases of oscillators.
Therefore, if the frequency mismatch |Δ| is less than some critical value Δcr =
|2ε| synchronization regime exists.
If the synchronization regime condition (4.8) becomes destroyed, e.g., due
to the increase of the frequency mismatch Δ, the phase difference is no more
bounded and in the original system (4.1) a beating regime appears. This regime
is characterized by the beating frequency Ωb which can be defined from (4.3)
for an arbitrary function q(θ) as:
2π −1
dθ
Ωb = 2π (4.9)
0 2εq(θ) − Δ
2.0
1.0
Ωb
0.0
−1.0
−2.0
−2.0 −1.0 0.0 1.0 2.0
Δ
Fig. 4.1. Beating frequency Ωb vs. frequency mismatch Δ for the coupling function
q(θ) = sin θ and ε = 0.5 in (4.3)
Conclusion
In this section we have shown that for weakly coupled oscillators the synchro-
nization problem can be solved only via the analysis of the phase dynamics
of the coupled systems. At that it is possible to say that the synchronization
regime appears for any large frequency mismatch (see conditions 4.8).
Mathematical Model
where we assume μ 1, α and β regulate the rate of the conservative and dis-
sipative coupling, respectively, γ and Δ define the amplitude and the frequency
mismatches. Averaging (4.12) over the period of possible synchronous oscilla-
tions we obtain:
4.1 Synchronization of Regular Systems 59
⎧
R12
⎪
⎪ Ṙ 1 = R 1 1 − β − + R2 (α sin(φ1 − φ2 ) + β cos(φ1 − φ2 )),
⎪
⎪ 4
⎪
⎪
⎨ Ṙ = R 1 + γ − β − R22 + R (−α sin(φ − φ ) + β cos(φ − φ )),
2 2 4 1 1 2 1 2
⎪
⎪ R1 φ˙1 = −αR1 + R2 (α cos(φ1 − φ2 ) − β sin(φ1 − φ2 )),
⎪
⎪
⎪
⎪
⎩
R2 φ˙2 = −(α + Δ)R2 + R1 (α cos(φ1 − φ2 ) + β sin(φ1 − φ2 )).
(4.14)
Writing this system in terms of the phase difference θ = φ1 − φ2 yields a
system that we are going to deal with in order to study synchronization in
the original system (4.12)
⎧
⎪ R2
⎪
⎪ Ṙ1 = R1 1 − β − 41 + R2 (α sin θ + β cos θ),
⎪
⎨
R2
Ṙ2 = R2 1 + γ − β − 42 + R1 (−α sin θ + β cos θ), (4.15)
⎪
⎪
⎪
⎪
⎩ θ̇ = Δ + α R2 − R1 cos θ − β R2 + R1 sin θ.
R1 R2 R1 R2
R1 = 2 + r1 ,
(4.17)
R2 = 2 + r2 ,
where r1 , r2 2. Then
r1 = α sin θ + β(cos θ − 1),
(4.18)
r2 = −α sin θ + β(cos θ − 1),
According to [143], (4.19) can have four, two, or no solutions at all. They
correspond to steady states of (4.15), and among them one or two may be
stable and these stable states refer to the synchronized regimes in the original
system (4.12). These synchronization solutions can be roughly classified into
two types:
1. In-phase if |θ| ≤ π/2
2. Antiphase (otherwise) synchronization regime
We illustrate the dynamics of the studied system (4.12) by plotting the
bifurcation diagrams in the (β, Δ)-plane for a sequence of α values (Fig. 4.2).
(a) (b)
7 1.4
6 α=1 OD 1.2
5
0S1
1
α = 0.3
4 B 0.8
Δ
α=1
Δ
3 0.6 α=0.5
α =0
2 0.4
α = 0.1
S
1 0.2 S2 α=0.33
0 0
−1 0 1 2 3 4 0 0.05 0.1 0.15 0.2 0.25 0.3
β α =0.27 β
Fig. 4.2. For the system (4.13): (a) Boundaries of the regions of the synchronization
regime (marked by S, for each value of α the region lies below the bifurcation curves).
The beating regime region is marked by B and the oscillator death region by OD.
(b) Boundaries of the bistability regime (for each value of α the bistability region
lies below the bifurcation curves and is marked by S 2 , the region of monostability
is marked by 0 S 1 (the left upper index refers to the in-phase regime: |θ| ≈ 0))
4.1 Synchronization of Regular Systems 61
The analysis of (4.15) and (4.16) allows to determine the frequency ωs and
the amplitude characteristics of the synchronized behaviors.
If α, β 1, then one can easily get the frequency of the synchronized
solution
Δ
ωs = −φ̇1 = − − α(1 − cos θ). (4.20)
2
By neglecting the terms of O(αβ) the frequency, observed in the system (4.12),
is ω̃s = 1 + μωs
2 .
According to (4.20), the in-phase solution implies that the synchronization
frequency ωs is equal to the mean frequency of both uncoupled oscillators:
Δ
ωs = −φ̇1 = −φ̇2 ≈ (4.21)
2
and that of the antiphase solution is
Δ
ωs = −φ̇1 = −φ̇2 ≈ 2α + . (4.22)
2
Numerical simulations yield that there is no qualitative difference in the
case when α and β are not small (Fig. 4.3a–d). This leads to the important
62 4 Synchronization of Two Coupled Systems
(a) (b)
1.8 −2.55
α =0.1 α= 0.27, Δ= 0.04
α =0.27 −2.6 α= 0.33, Δ= 0.04
1.6 α =0.33 α= 0.5, Δ=0.2
α =0.5 −2.65 α = 1, Δ= 0.8
1.4
−2.7
1.2 −2.75
φ
−2.8
φ
1
−2.85
0.8 −2.9
0.6 −2.95
−3
0.4
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 −3.05
0 0.05 0.1 0.15 0.2 0.25
β β
(c) (d)
0.515 2.5
α =0.1
0.51 α=0.27, Δ= 0.04
α =0.27
α =0.33 α=0.33, Δ= 0.04
0.505 α =0.5 α=0.5, Δ =0.2
2 α= 1, Δ =0.8
0.5
0.495
ωs
0.49
ωs
1.5
0.485
0.48
0.475 1
0.47
0.465
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.5
0 0.05 0.1 0.15 0.2 0.25
β β
(e) (f)
2.2
2.1
2
2 1.9
1.8
R1,2
R1,2
1.7
1.6
1.5
1.5
1.4
1.3
1.2
1 0 0.05 0.1 0.15 0.2
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
β β
Fig. 4.3. Phase difference θ̄, synchronization frequency ωs , and amplitude char-
acteristics R̄1,2 of the in- and antiphase solutions of (4.13) in dependence on the
dissipative coupling β. (a, c) stable in-phase solution for fixed Δ = 1, (b, d) stable
antiphase solution, typical amplitude values for (e) in-phase (α = 0.5, Δ = 1) and
(f ) antiphase (α = 0.5, Δ = 0.2) synchronization regimes (R1 solid lines, R2 dashed
lines)
Frequency asymmetry. Now let us turn to the area between the beating
regime and the oscillator death regions with a characteristic value of β near 1
(Fig. 4.2a) that was described before as an optimum coupling value. The corre-
spondent boundary curves have an asymptote β = 1; the width of this region
decreases to zero for Δ → ∞. Increasing Δ gradually, we find numerically
that the synchronization properties change dramatically. The synchronization
frequency deviates from the mean frequency and approaches that of the first
oscillator, i.e.,
ωs = α. (4.23)
Hard and soft transitions to the synchronous regime. As mentioned in
Sect. 4.1.1, another important characteristic of interacting systems, where syn-
chronization can take place, is the beating frequency Ωb .
Three types of qualitatively different oscillatory behavior may be observed
in the averaged system (4.13):
• θ = φ1 − φ2 = const corresponds to the synchronization regime, Ωb = 0.
Regime of strong synchronization takes place
• |θ| = |φ1 − φ2 | ≤ const means that a stable limit cycle exists, and changes
in the phase difference of the solution are limited. This is called “phase
entrainment” [155] or “phase trapping” [145]; Ωb is still zero, i.e., this is a
synchronization regime as well, but this regime is nonstrong
• Finally, if |φ| = |φ1 − φ2 | grows unbounded, Ωb = 0 now, then a stable
limit cycle (also called a “libration orbit” [155]) corresponds to a phase
drift regime. Thus, in the original system (4.12) a beating regime takes
place.
The transition from the second to the third type of behavior was the
subject of study in [155] for the special case α = 0, Δ = 0, β = γ = 0.
Next, we describe the routes of losing the synchronization regime appearing
in this system. By numerically performed simulations of the original system
(4.12) for μ = 0.1, the averaged beating frequency Ωb is calculated for γ = 0,
varying Δ, a fixed α, and a sequence of β values (Fig. 4.4a) and vice versa
(Fig. 4.4b). Indeed, when a stable rest state disappears through a saddle-node
bifurcation, a stable limit cycle with a phase difference growing unbounded is
born instead. There are two kinds of transitions (1) As long as the flow slows
down near the location of the former rest state, the period of this cycle is
very large near the bifurcation point, making the averaged beating frequency
increasing continuously (Fig. 4.4a), i.e., a soft transition to (from) synchro-
nization regime takes place. Nevertheless, the larger β, the sharper the slope
of the frequency curve is. (2) A quite opposite scenery is observed, when an
Andronov–Hopf bifurcation occurs. A limit cycle with a bounded variation
64 4 Synchronization of Two Coupled Systems
(a) 2.5
(b)
2.5
β=0.1 α=0.1
β=0.3 α=0.3
2 β=0.5 2 α=0.5
β=0.7 α=0.7
β=0.9 α=0.9
1.5 1.5
Ωb
1
Ωb
0.5 0.5
0 0
0.5 −0.5
0 1 2 3 4 5 0 1 2 3 4 5
Δ Δ
Fig. 4.4. Beating frequency Ωb , calculated for the original system (4.12) for μ = 0.1
and γ = 0, illustrates soft and hard transitions from the synchronous to the beating
regime. (a) fixed α = 0, (b) fixed β = 0.5
(b)
25
(a) 1.5
20
1 B
15
OD
Δ
Δ
B 0S1
10
0.5
5 S2
S
0 0
0 0.5 1 1.5 2 2.5 3 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
β β
25
(c) 1.5 (d)
20 B
1
15
S
Δ
πS1
Δ
B OD 0S 1
10
0.5
5
S S2
0 0
0 0.5 1 1.5 2 2.5 3 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
β β
20 (f) 1.5
(e)
18
16
B
14
B OD 1
12 S
10
Δ
Δ
8
0.5 S1
6
4
2
0 0
0 0.5 1 1.5 2 2.5 3 3.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
β β
Fig. 4.5. Influence of the amplitude mismatch γ in the system (4.13). (i) Regions
of synchronization (S 1 ) for α = 0.5 and (a) γ = 0, (c) γ = 1, (f ) γ = 2; (ii) regions
of bistability (S 2 ), surrounded by regions of in- (0 S 1 ) and antiphase (π S 1 ) synchro-
nization regime. α = 0.5 and (b) γ = 0, (d) γ = 0.1, (e) γ = 1.2
66 4 Synchronization of Two Coupled Systems
Conclusions
Δ
(a) (b) (c)
1.0
0.8 -
0.6 -
0.4 -
0.2 -
0.0
0. 1. 0. 1. 0. 1.
d d d
Fig. 4.6. Phase diagrams of different synchronization regimes in (4.24) are presented
on the plane (Δ = ω2 − ω1 , d). Parameters are: ω1 = 1, a = 0.5 (a), a = 0.8 (b), and
a = 0.95 (c). Synchronization regions are indicated in gray colors, beating regions
are white. In all diagrams the main gray region corresponds to 1 : 1 synchronization
regime. The gray regions to the left and to the top correspond to 2 : 1, 3 : 1, 4 : 1,
etc. synchronization regimes
68 4 Synchronization of Two Coupled Systems
Conclusions
The main finding of this section is that nonuniformity of rotations can lead
to various n : m synchronization while for the uniform rotation only 1 : 1
synchronization is possible.
where d is the coupling strength. ω1,2 determine the mean frequency of the
oscillators in the case of phase-coherent attractors. In our study we take f =
0.1, ω1 = 0.98, and ω2 = 1.02. The parameter a is changed in the interval
[0.15; 0.3].
As already mentioned in Sect. 2.3, for phase-coherent as well as funnel
attractors occurring in the Rössler oscillator, projections of chaotic trajecto-
ries onto the plane (ẋ, ẏ) always rotate around the origin and therefore the
phase can be defined as3
ẏ
φ = arctan , (4.26)
ẋ
while the instantaneous frequency is:
ẏẍ − ÿ ẋ
ν= . (4.27)
ẋ2 + ẏ 2
We then use again two criteria to detect the existence of CPS4 :
1. Locking of the mean frequencies
Ω1 = ν1 = Ω2 = ν2 , (4.28)
0.20
l1
0.15 l2
l3
0.10
d
0.05
0.00
0.15 0.20 0.25 0.30
a
Fig. 4.7. Critical coupling curves of the Rössler systems (4.25). l1 corresponds to
the onset of CPS, i.e., below this line the oscillations are not synchronized, and above
both the phase- (4.29) and frequency- (4.28) locking conditions are fulfilled; l2 to the
transition of one of zero LEs to negative values and l3 to zero crossing of one of the
positive LEs. Note: In this figure we do not separate cases, where synchronization
occurs between regular and chaotic oscillations
Strong Coherence
Let us start with the simplest case in the interval [0.15; 0.186], where both
oscillators have phase-coherent chaotic attractors (see Figs. 2.3a and 2.4a, d).
For these attractors the zero LEs are associated with the phase dynamics.
Due to the high degree of coherence of the motions, i.e., a very small diffusion
constant Dφ (see Fig. 2.3), phase and frequency locking occur shortly after the
transition of one of the zero LEs to a negative value. Note that the two largest
LEs remain positive, i.e., hyperchaos remains. Hence, the amplitudes of the
oscillators are only weakly correlated. A strong correlation of the amplitudes
sets in only at a much larger coupling (d > dl3 ) where one of the positive
LEs becomes negative, and the two systems achieve a special type of almost
complete synchronization (CS).5 Between phase and complete synchronization
usually it is possible to observe lag synchronization [166], where the motions
of two chaotic systems are nearly identical, but one oscillator lags in time to
the other.
PS of phase-coherent chaotic Rössler oscillators can be better understood
by separating the original system into the dynamics of amplitude and phase.
By introducing6
5
Complete synchronization (x1 (t) = x2 (t), y1 (t) = y2 (t), z1 (t) = z2 (t)) is impossi-
ble because the oscillators are nonidentical. But in the following for the shortness
we will call this regime “complete synchronization.” In this regime with increase
of coupling the states of two (or more) systems becomes closer and closer.
6
This approach is valid only for the phase-coherent case.
4.2 Synchronization of Coupled Chaotic Oscillators 71
we get
In this system phases are the slowest variables in comparison to the other
variables ρ1,2 and z1,2 . Therefore we can use averaging over rotations of the
phases. Introducing the phase difference θ = ψ1 − ψ2 of the “slow” phases
ψ1,2 according to φ1,2 = ω0 t + ψ1,2 , the frequency difference Δ = ω1 − ω2 and
averaging the equations for them, one gets
dθ d ρ2 ρ1
= Δ − ( + ) sin θ. (4.32)
dt 2 ρ1 ρ2
When we neglect the fluctuations of the amplitudes, (4.32) is reduced to the
Adler equation which has the stable steady-state solution
2Δρ1 ρ2
θ = arcsin (4.33)
d(ρ21 + ρ22 )
when the coupling strength d is larger than the critical value
dps = Δ (4.35)
dps is then the onset of PS for phase-coherent attractors. This makes clear that
CPS is very similar to the classical case of phase synchronization of coupled
periodic oscillators (see Sect. 4.1).
Intermediate Coherence
A quite different route to CPS takes place for a ∈ [0.195, 0.25] where both
chaotic attractors are of the funnel type (see Figs. 2.3b and 2.4b, e). Here
the curves l1 and l2 in Fig. 4.7 are clearly separated, but both lie below the
curve l3 . Hence, the two largest LEs remain positive during the transition
to CPS, which means that no bifurcation of the hyperchaotic attractor can
be associated to the locking of the phases. CPS occurs here via a crisis-like
transition inside the hyperchaos, i.e., via an interior crises – strong widening
(or narrowing) of a preexisting chaotic attractor [70] of the chaotic set. Such
a change of the internal structure of the attractor is seen by the projection
of the phase trajectory on the plane (φ1 , φ2 ) for d outside (Fig. 4.8a) and
72 4 Synchronization of Two Coupled Systems
φ1 φ1
Fig. 4.8. Projections of trajectories of the Rössler systems (4.25) for a = 0.22 on
the plane (φ1 , φ2 ) for the coupling strength d outside (a) (d = 0.055) and within
(b) (d = 0.075) the synchronization region
Strong Noncoherence
In the interval a ∈ [0.25, 0.3], the curve l1 lies above the curve l3 , showing
that CPS occurs only after one of the positive LEs passes to a negative value,
i.e., the transition to CS. It is important to note that, by CS, both oscillators
have established a rather strong cross-correlation. However, such a strong
relationship is an average property over the whole attractors, while locally,
phase slips associated to different number of oscillations in the two oscillators
in a period of time may occur for coupling strengths d shortly after l3 , as
seen in Fig. 4.9 as the typical behavior for d ∈ [dl3 , dl1 ]. CPS appears now
as a manifestation of CS. This property is in contrast to the above regimes,
where CPS is a much weaker degree of synchronization compared to CS, and
the phases become locked before a strong correlation of the amplitudes is
established. Thus, for highly noncoherent oscillations due to the existence of
two distinct characteristic time scales, a rather strong coupling is necessary
to keep both oscillators in small or larger cycles simultaneously in order to
maintain the phase locking. Otherwise, a phase slip develops quickly due to the
very different time scales when the two oscillators are on a small and a large
4.2 Synchronization of Coupled Chaotic Oscillators 73
10
(a)
6
φ2−φ1
−2
20
(b)
10
dy1/dt, dy2/dt
−10
−20
27100 27150 27200 27250 27300 27350
time
Fig. 4.9. CPS in the strong noncoherence regime. (a) Time evolution of phase
difference. (b) Variables ẏ1,2 in system (4.25) for a = 0.2925 and d = 0.179 ∈
[dl3 , dl1 ]. Solid and dotted lines correspond to the first and the second oscillator,
respectively. In the time interval between dashed lines the first oscillator produces
four rotations in the (ẋ1 , ẏ1 )-plane around the origin, but the second one generates
only three rotations, which leads to a phase slip in (a)
1.08
a=0.16
1.05
a=0.22
Ω2/Ω1
a=0.28
1.02
0.99
0.00 0.05 0.10 0.15 0.20
d
Fig. 4.10. Mean frequency ratio Ω2 /Ω1 vs. coupling d for the Rössler systems (4.25).
The fitting curve (∼ |d − dl1 |0.5 ) for a = 0.16 (dashed line) and fitting straight lines
for a = 0.22 (solid line) and a = 0.28 (dotted line) are presented
the trajectory rotations. When the coupling is weak, it appears very often that
the first oscillator is on the small loops, while the second one is on the large
ones, or vice versa. This leads to an unpredictable, oscillatory-like evolution of
the Ω2 /Ω1 -ratio. Relatively close to the critical point, rather strong coupling
makes both oscillators stay simultaneously on small or large loops for a long
period of time. The Ω2 /Ω1 -ratio decreases monotonous and can be fitted by
a straight line in a large range below the critical point.
The difference in the transitions to CPS may also be understood by means
of UPOs [119]. It was shown that CPS takes place in a parameter region where
all pairs of UPOs, embedded in the chaotic attractors, are synchronized. For
phase-coherent attractors the period of all UPOs are close to each other (see
Fig. 2.3g) and the boundaries of the synchronization regions in the form of
Arnold tongues corresponding to the locking of different pairs of UPOs lie in
a relatively narrow region. The approach to CPS, thus, is associated to an
effective saddle-node bifurcation with small noise and a type-I intermittency
occurs, resulting in a relatively hard transition to CPS. For the funnel chaotic
attractors the distribution of the mean periods of UPOs is rather broad (see
Fig. 2.3h, i). Therefore, the coupling strengths corresponding to the onset of
synchronization of different pairs of UPOs are distributed in a rather large
interval. With an increase of d, the system crosses continuously a series of well-
separated Arnold tongues, resulting in a much slower convergence to CPS. The
scaling lows presented in Fig. 4.10 support these treatment.
0.4
OD
0.3
NS
Δ
0.2
S
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
d
Fig. 4.11. The regions of nonsynchronous (NS) and synchronous (S) motion and
absence of oscillations (oscillation death) (OD) in system (4.25). The diagram is
approximate, e.g., windows of periodic behavior in regions I and II are not shown.
Compare with the phase diagram in Fig. 4.2 for the van der Pol oscillators
Note that the boundaries between these different regimes are slightly dis-
persed, and windows of periodical behavior are present, although for a large
domain of parameters in the regimes I and II the amplitudes of the oscillations
are chaotic.
Next, we discuss the behavior of the beating frequency Ωb , when the fre-
quency mismatch Δ changes (Fig. 4.12). We see that for weak coupling (small
d), the beating frequency Ωb smoothly depends on Δ. This means that a
regime, where the frequencies of interacting oscillators differ by a rather small
value, is possible. For sufficiently strong coupling the transition from synchro-
nous to a nonsynchronous state is rather sharp: a virtual jump in dependence
76 4 Synchronization of Two Coupled Systems
0.20
d=0.002
0.05
0.1
0.15 0.14
Ωb
0.10
0.05
0.00
0.00 0.05 0.10 0.15 0.20
Δ
Fig. 4.12. Dependence of the beating frequency Ωb on the frequency mismatch Δ
for different values of coupling d in (4.25). In full analogy to the classical case of
two coupled periodic oscillators, the transition to (or from) synchronization regime
occurs smoothly or practically by a jump for weak and strong coupling, respectively
Conclusions
In Sect. 3.5 the effect of external CPS of intermittent chaotic oscillators due
to a periodic driving was demonstrated. As the examples of such oscillators
the Lorenz system and quadratic map have been used. Now we will show that
the mutual synchronization of such systems is possible as well. Because both
mentioned systems have their well-pronounced characteristic time scale, with
increase of coupling, it is possible to expect the coincidence of averaged length
of laminar stages, which means the existence of synchronous regimes.
Synchronization of Maps
ẋ1 = σ(y1 − x1 ),
ẏ1 = r1 x1 − y1 − x1 z1 + d(y2 − y1 ),
ż1 = −bz1 + x1 y1 ,
(4.38)
ẋ2 = σ(y2 − x2 ),
ẏ2 = r2 x2 − y1 − x2 z2 + d(y1 − y2 ),
ż2 = −bz2 + x2 y2 ,
1580
ε1=4*10−6
ε2=4.5*10−6
1560
1540
<τ 1,2>
1520
1500
1480
0 0.0001 0.0002
d
0.25
0.20
0.15
0.10
0.05
xn, yn
0.00
−0.05
−0.10
−0.15
−0.20
−0.25
50000 55000 60000 65000 70000
time
Fig. 4.14. Time series for synchronized maps (4.36) and (4.37). xn (solid line) and
yn (dashed line). Parameters are: g = 2, ε1 = 4 ∗ 10−6 , ε2 = 4.5 ∗ 10−6 , d = 0.00015
0.16
1/t1
1/t2
0.14
0.12
0.1
1/ti
0.08
0.06
0.04
0.02
0
10−4 10−3 10−2 10−1 100 101
d
2
(a)
0.02 (b)
1.5
λ
y
0
0.5
0
−3 −1.5 0 1.5 3 0 0.005 0.01 0.015
φ d1
300 (d)
1
d1=0.0065 (c)
200
d1=0.007
φ2−φ1
Ω1/Ω2
100
d1=0.0072
0.995
d1=0.008
0
0 50000 100000 150000 200000 0 0.005 0.01 0.015
time d1
Fig. 4.16. Synchronization of rotatory phase variables. (a) Projections of the typical
rotatory trajectory of the uncoupled system in (4.39) on the plane (φ, y). Parameters
are: γ = 0.645, μ = 3.0. In Fig. 4.16b–d parameters are: γ1 = 0.645, γ2 = 0.667,
μ = 3.0, and d2 = 0. (b) The four largest LEs, one of which is always zero.
(c) Difference of phase variables φ2 − φ1 for nonsynchronous (d1 = 0.0065; 0.007;
0.0072) and synchronous (d1 = 0.008) regimes. (d) The mean frequency ratio Ω1 /Ω2
vs. coupling
1 0.04
(a)
(b)
0.03
0.5
0.02
λ
y
0.01
0
0
−0.5 −0.01
0 1 2 3 0 0.005 0.01 0.015
φ d
600 1.01
(c) (d)
d=0.006
400
d=0.007
Ω1/Ω2
ψ1−ψ2
200
d=0.008
0
d=0.009 1
0 50000 100000 150000 200000 0 0.005 0.01 0.015
time d
synchronous state. But usually [173, 174] the transition to CS is very close to
the point when one of the positive LEs becomes negative. Therefore, we will
infer the onset of CS from this property.
Due to the second property, i.e., chaos in continuous in time system pos-
sesses a zero LE, there are many properties in common between phase synchro-
nization of autonomous chaotic oscillators (see previous sections) and phase
synchronization of autonomous chaotic phase systems.
As for periodic synchronization, the appearance of CPS is affected by the
frequency mismatch of the coupled subsystems and by the coherence property
of the motions. We will characterize this coherence, i.e., the diffusion of the
phase variables, by their variances Dφ1,2 that are defined for large times as
We will show below that these Dφ1,2 of both coupled subsystems, as well as
their frequency mismatch, play the crucial role in the transitions to phase
synchronization regime.
82 4 Synchronization of Two Coupled Systems
In this case the phase variables φ1,2 unboundedly increase and φ̇1,2 are always
(or almost always) positive. A projection of the chaotic phase rotating trajec-
tory on the plane (φ, y) (Fig. 4.16a) looks like a “smeared” periodic trajectory
with a monotonously (or almost always) increasing phase. Therefore, PS of
chaotic rotations is quite similar to periodic synchronization, i.e., in both
cases only the phase growth rate is important. The averaged growth rate of
the phases or the mean frequency of rotations can be defined as
Ω = φ̇ = y . (4.41)
this large difference is that PS takes place via a crisis transition of the struc-
ture of the hyperchaotic attractor, i.e., via an interior crises of the chaotic set.
The corresponding phase portraits look very similar to the phase portraits in
Fig. 4.8 obtained for coupled chaotic Rössler oscillators.
At larger coupling (d21 ≈ 0.0118), where one of the positive LEs becomes
negative, CCPS occurs. Due to the relatively high noncoherence properties,
the interval of values of coupling between the transitions to RCPS and to
CCPS L = [d11 ; d21 ] is small. As our numerical simulations show, the increase
of the parameters γ1,2 leads to a complication of the topological structure
of the chaotic attractors. The intervals, where the phase variables decrease,
become larger and the behavior transfers from a rotational to an oscillation-
rotational one. This leads to an increase of the noncoherence properties of
motion (diffusion of the phase variable increases) and as a result the width of
the interval L between RCPS and CCPS tends to zero. The reason for that
is the following. CPS is quite similar to the synchronization of periodic oscil-
lations in the presence of small noise [1, 163]. When noise increases, a larger
coupling is needed to achieve phase locking. By analogy, in order to suppress
large phase fluctuations by CPS, a stronger coupling has to be applied.
In this case in both subsystems in (4.39) the phase variable oscillates around
some constant value, i.e., φ1,2 are bounded (Fig. 4.17a). Synchronization of
such oscillatory phase variables is quite similar to the case of usual phase
synchronization of chaotic oscillators (see Sect. 4.2.1). Because of the simple
topology of the chaotic attractor, we can introduce a new “artificial” phase:
y
ψ = arctan , (4.44)
φ − arcsin γ
a new amplitude
A = ((φ − arcsin γ)2 + y 2 )1/2 (4.45)
and this yields the mean frequency
ψ(T ) − ψ(0)
Ω = ψ̇ = lim . (4.46)
T →∞ T
Here the conditions (4.42) and (4.43) applied to the new phase variables
ψ1,2 and the mean frequencies Ω1,2 can be used as criteria for synchroniza-
tion regime. Therefore, although the oscillatory and rotatory cases cannot be
generally reduced one to the other, two similar criteria for the existence of
phase synchronization regime can be used and as we will show, many sim-
ilar effects happen. For the chosen parameters: γ1 = 0.815, γ2 = 0.83, and
μ1,2 = 3.3, the coherence of the motions is significantly stronger compare to
the rotatory case (Dψ1 ≈ 0.075, Dψ2 ≈ 0.079). We consider y- and z-coupling
(in (4.39) d1 = d2 = d). As in the case of phase synchronization of the
84 4 Synchronization of Two Coupled Systems
2.5 2.5
(a) (b)
2 2
1.5 1.5
φ2
φ2
1 1
0.5 0.5
0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
φ1 φ1
Fig. 4.18. Projections of the trajectories of the system (4.39) on the plane (φ1 , φ2 )
outside of the synchronization region – (a) (d = 0.008) and within the synchro-
nization region – (b) (d = 0.009). The parameters are: γ1 = 0.815, γ2 = 0.83, and
μ1,2 = 3.3
4.2 Synchronization of Coupled Chaotic Oscillators 85
2
(a)
0.02 (b)
1
0
0
λ
y
−0.02
−1
−0.04
−2 −0.06
−3 −1.5 0 1.5 3 0 0.005 0.01 0.015
φ d
300
(c) (d)
d=0.007 1.005
200
φ1−φ2
Ω1/Ω2
d=0.0078
100
d=0.008
1
0
0 50000 100000 150000 200000 0 0.005 0.01 0.015
time d
0.02
Ω1−Ω2, λ
0.01
−0.01
0 0.005 0.01 0.015
d
Fig. 4.20. Synchronization of oscillatory–rotatory phase variables. CS occurs before
CCPS. The parameters are: γ1 = 0.34, γ2 = 0.39, μ1,2 = 5.0, and d1 = d2 = d in
(4.39). The three largest LEs and the mean frequency difference Ω1 − Ω2 (circles)
vs. coupling are given
4.2 Synchronization of Coupled Chaotic Oscillators 87
0.03 1.01
(a) (b)
0.02
Ω2/Ω1
0.01
λ
−0.01
1
−0.02
0 0.005 0.01 0.015 0 0.005 0.01 0.015
d1 d1
(c)
d1=0.0
1000
d1=0.008
φ2−φ1
500
d1=0.0084
d1=0.0088
0
0 50000 100000 150000 20000
time
Fig. 4.21. Hard transition to RCPS in (4.39). The parameters are: γ1 = 0.645,
γ2 = 0.636, μ1 = 3.0, μ2 = 3.05, and d2 = 0. (a) The four largest LEs. (b) The mean
frequency ratio Ω1 /Ω2 vs. coupling. (c) Difference of the phase variables φ2 − φ1 for
nonsynchronous (d1 = 0.0; 0.008; 0.0084) and synchronous (d1 = 0.0088) regimes.
The diffusion constants are Dφ1 ≈ 0.219 and Dφ2 ≈ 0.218
88 4 Synchronization of Two Coupled Systems
Fig. 4.22. Projections of the trajectories of system (4.39) on the planes (φ1 , φ2 ) ((a)
and (b)) and (y1 , y2 ) ((c) and (d)) for γ1 = 0.645, γ2 = 0.636, μ1 = 3.0, μ2 = 3.05,
and d2 = 0 outside of the synchronization region – ((a) and (c)) (d1 = 0.0084) and
within the synchronization region – ((b) and (d)) (d1 = 0.0088)
The relatively large jump in the mean frequency ratio ρ = Ω2 /Ω1 from a
nonsynchronous (ρ = 1) to a synchronous (ρ = 1) hyperchaotic behavior at
d1 = 0.0088 can be regarded as a manifestation of a hard transition to phase
synchronization regime. Indeed, for very small changes in the coupling, strong
changes in the phase difference evolution (Fig. 4.21c) and in the phase portrait
(Fig. 4.22) are observed. For d1 = 0.0084, i.e., when d1 is very close to the crit-
ical value d11 , only very short intervals of synchronized epochs are observed in
the phase difference (compare with Figs. 4.16c, 4.17c, and 4.19c) that demon-
strate phase differences for the oscillatory case where the transition to PS is
hard. The projections of the hyperchaotic attractor on the planes (φ1 , φ2 ) and
(y1 , y2 ) slightly before and after the transition to phase synchronization regime
are presented in Fig. 4.22. For the synchronous regime the chaotic trajectory
lies within relatively narrow bands in the phase space (Fig. 4.22b, d), while
when synchronization regime is lost these bands smear and merge together
(Fig. 4.22a, c). Such a hard transition to a band-structured attractor can be
explained in terms of unstable periodic orbits. In [117] it was shown and in
previous sections described in detail that CPS regime occurs in that parameter
region, where all unstable periodic orbits embedded in the chaotic attractors
are synchronized. For the presented case the hard transition to phase synchro-
nization regime is caused by the fact that boundaries of the Arnold tongues,
4.2 Synchronization of Coupled Chaotic Oscillators 89
Conclusions
1. Real chaotic phase synchronization (RCPS), which occurs when two LEs
are positive and when both the frequency- and the phase locking condi-
tions are fulfilled
2. Complete chaotic phase synchronization (CCPS), which occurs when only
one LE is positive and when both the frequency- and the phase locking
conditions are fulfilled
3. Complete chaotic synchronization (CS), which occurs when only one LE
is positive and when the frequency- and the phase locking conditions are
both not fulfilled.
We begin our analysis with the simplest case of coupled shift maps (c = 0),
i.e., system (4.47) reads now:
φk+1
1 = ω1 + φk1 + d sin (φk2 − φk1 ),
(4.50)
k+1
φ2 = ω2 + φk2 + d sin (φk1 − φk2 ).
Noncoherent Case (c = 0)
2.5 dcr
*
d
2.0
1.5
Δ
1.0
1:1 synchronization
0.5
0.0
0.0 0.5 1.0 1.5 2.0 2.5
d
Fig. 4.23. Critical values of coupling d∗∗ and d∗ corresponding (i) to the transition
from nonsynchronous to synchronous motion (left curve) and (ii) to the transition
from synchronous to nonsynchronous motion (right curve) in the model (4.50) vs.
the frequency mismatch Δ. Between both curves there is the region of 1:1 synchro-
nization
1.5 0.7
(a) (b)
0.6
1.4
0.5
1.3
0.4
w
0.3
1.2
c=0,0 c=0.0
0.1 0.2
0.2 0.4
0.4 0.2
0.6
0.6 0.8
1.1 0.9 1.2
1.0 1.6
0.1
1.0 0.0
0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3 0.4 0.5
d d
Fig. 4.24. The winding number w vs. the coupling coefficient d at different c and
ω1 = π/2, ω2 = 2π/3 (a), respectively, ω1 = 4π/7, ω2 = 17π/23 (b) in system (4.47)
We now carry out an analysis of synchronization for two coupled chaotic CMs
(c < 0). In this system some different synchronization properties are observed.
We perform numerical simulations for fixed ω1 = 0.6 and different values of ω2
(Fig. 4.25). Usually there exists only the region of 1:1 synchronization regime.
Only in rather small intervals of c, regions of m : n synchronization regime
can occur. It should be noted that in all our presented experiments only m : 1
synchronization regimes with different m = 2, 3, 47 are observed. Figure 4.25
indicates that the geometrical structure and the sizes of the synchronization
7
If we take larger value of ω2 , a synchronization regime with larger m takes place
too.
94 4 Synchronization of Two Coupled Systems
−3.0
(a) (b) (c) (d) (e) (f) (g) (h) (i) (j)
−2.5 -
−2.0 -
−1.5 -
−1.0 -
−0.5 -
c=0 +
0. 1. 2.
d d d d d d d d d d
Fig. 4.25. Regions of chaotic phase synchronization for ω1 = 0.6 and different
values of ω2 : 0.8 (a), 1.0 (b), 1.2 (c), 1.4 (d), 1.6 (e), 1.8 (f ), 2.0 (g), 2.2 (h), 2.4 (i),
2.6 (j) in (4.47). The main gray regions correspond to 1:1 synchronization regime.
In columns (c–j) for relatively small −c small regions of 2:1 (c–g), 3:1 (f –h) and
4:1 (i, j) synchronization are presented. They are visible as small stripes in the left
bottom areas
2. Large −c for which the variance D is very large (see Fig. 2.13) and due
to a high noncoherence of the motions, synchronization regime cannot be
achieved; in our simulation this is the interval D3 : c < −2.25
3. Intermediate c that do not belong to the two previously defined intervals;
this is the interval D2 : c ∈ [−0.6; −2.25]
For each of these intervals we now analyze the influence of the three para-
meters on synchronization discussed in Sect. 2.5.4, i.e., the variance D, the
parameter γ characterizing the relative duration of intervals of phase increase
and phase decrease, and the rotation number difference Δρ = ρ2 − ρ1 .
In the interval D1 the difference of the rotation numbers Δρ plays the cru-
cial role in getting synchronization. As in the case of regular coherent CMs
(Sect. 4.3.1), the critical value of coupling d+ , at which the transition to 1:1
synchronization regime occurs, depends on the value of the rotation number
difference Δρ. At larger values Δρ, a larger value d is needed to achieve syn-
chronization. The sizes of the synchronization regions become smaller with
increase of c. This happens due to the increase of the noncoherence of the
rotation. At chosen values ω1 and ω2 in the interval D3 , synchronization is
in general impossible due to the highly noncoherent properties of rotations.
For instance, at c = −2.5 and ω1 = 0.6 resp. ω2 = 2.6, we find that imperfect
PS (i.e., seldom occurring phase slips are possible, cf. Sect. 3.3) exists at a
very small frequency mismatch Δ = 0.0001. Therefore, a very small Δ and
as a result of that a very small rotation number difference (remember that
it is the average characteristic) does not always guarantee the occurrence of
a phase synchronization regime, because the complexity, specifically the non-
coherence, of the behavior quantified by the variance D of rotations can be
crucial. The existence of time intervals with a strongly different phase growth
rate, which leads to the absence of well-pronounced characteristic time scale,
makes the processes of locking of rotations impossible.
Intermediate Noncoherence
−c > ω. So for the first element with ω1 = 0.6, this critical value is equal
to −0.6. Figure 2.14 indicates that γ (the ratio of the duration of the phase
decreasing intervals to the duration of phase increasing intervals for the first
element) becomes strongly increasing at c ≈ −1. If in the second element the
phase is still monotonically increasing, then the time interval, where the phases
rotate in opposite directions, are existing for coupled elements. This makes
the phase entrainment rather difficult and usually a phase synchronization
regime does not exist.8 If with an increase of c in the second element phase
decreasing intervals appear, rotations in both elements become more similar,
i.e., in both elements the phases can grow and vanish, and a phase entrainment
can happen. We expect that this mechanism leads to the existence of islands of
synchronous motions for several values of frequency mismatch (Fig. 4.25c–g),
because these islands appear at such values c that approximately correspond
to the transition to the second type of rotations in the second CM.
Synchronized Hyperchaos
λ1 = log | 1 − c
π |,
(4.52)
1
M
λ2 = limM →∞ M k=1 log | 1 − c
π − 2 d cos(φk2 − φk1 ) |.
Since the first LE λ1 is constant and positive for all values of d, we infer that
only the sign of the second LE λ2 is important for the occurrence of chaotic
PS. If both LEs are positive, there is a hyperchaotic regime that determines
usually a nonsynchronized regime. If, with increase of coupling, λ2 becomes
negative, there is a strong indication for the occurrence of CPS. This situation
takes place at the transitions to 1:1 synchronization regime in all simulations
presented in Figs. 4.26 and 4.27.
But this is not the only scenario for the transition from nonsynchronous
to synchronous behavior for which criteria (4.48) and (4.49) are satisfied.
This is illustrated by the plots showing the dependencies of the winding
number w = ρ2 /ρ1 and the second LE λ2 on the coupling coefficient d
(Fig. 4.26) and phase diagrams for nonsynchronous (Fig. 4.27a, b) and syn-
chronous (Fig. 4.27c) regimes. In the interval d ∈ [0.285, 0.32] the winding
number w = 3/1 that corresponds to a 3:1 synchronization regime, but the
second LE remains positive λ2 ≈ 0.05, i.e., synchronized hyperchaos exists.
Also there are intervals of d in which 2:1 and 1:1 hyperchaos synchronizations
8
It is interesting to note that as one can see from the conditions for the existence
and stability of a fixed point in the sine circle map, the 1:1 synchronization can
occur for CM with contrary rotating phases, e.g., for ω1 < 0 < ω2 .
4.3 Synchronization of Coupled Circle Maps 97
4.0
w
3.0
λ2
2.0
1.0
0.0
−1.0
0.10 3.1
−2.0
−3.0
0.05 3.0
−4.0
−5.0
0.00 2.9
0.25 0.30 0.35 0.25 0.30 0.35
−6.0
0.0 0.5 1.0 1.5 2.0
d
Fig. 4.26. The winding number w = ρ2 /ρ1 and the second LE λ2 vs. the coupling
coefficient d for ω1 = 0.6, ω2 = 2.0, and c = −0.15 in (4.47). Regions of 3:1, 2:1, and
1:1 synchronization exist. Enlargements of the interval [0.25;0.35] are presented in
the insets (left: λ2 , right: w)
Fig. 4.27. Phase portraits of system (4.47) for ω1 = 0.6 , ω2 = 2.0, c = −0.15 and
different d within (c) (d = 0.3) and outside (a) (d = 0.25) and (b) (d = 0.275) of the
3:1 synchronization region. In all three cases the system (4.47) is in a hyperchaotic
regime (λ1 , λ2 > 0)
30
d = 0.28
20
d = 0.282
10
φ2k−3φ1k
d = 0.3
0
−10 d = 0.322
d = 0.326
−20
−30
0 2000 4000 6000 8000 10000
k
Fig. 4.28. Evolution of the phase difference θk = φk2 −3φk1 for synchronous (d = 0.3)
and nonsynchronous (d = 0.28; 0.282; 0.322; 0.326) regimes of system (4.47). The
parameters are: ω1 = 0.6, ω2 = 2.0, and c = −0.15
dense bands of motions can be clearly seen (Fig. 4.27b, c). From the synchro-
nization point of view, the attendances of gaps between these bands are corre-
sponding to slips in the phase difference θk = φ2 − 3φ1 , i.e., jumps of 2π [117]
(see Fig. 4.28). A decrease of the number of slips exhibits the tendency of the
system to perfect PS where no slips exist. At synchronized hyperchaos, the
chaotic trajectory is concentrated only in relatively narrow bands in the phase
space (Fig. 4.27c). This transition to synchronous motions corresponds to the
transition of the phase difference θk = φk2 − 3φk1 from rotation to oscillation.
Thus the transition from nonsynchronous to synchronous behavior in a two-
element CMs system occurs through an interior crisis [79] of the hyperchaotic
set, i.e., in both regimes both LEs are positive.
Conclusions
(a) Due to the discreteness the systems can have equal time scale charac-
teristics (rotation numbers) for different parameters;
(b) The increase of coupling leads to the destruction of synchronization
regime, i.e., there exists an interval of coupling for which a synchro-
nization regime takes place;
(c) Strong noncoherence of motions is a reason for nonpredictable
synchronization transitions.
Part II
In this chapter we are starting with the main part of this book, the treatment
of synchronization phenomena in ensembles or networks of coupled oscilla-
tors. First we treat networks of coupled first-order phase oscillators. This
choice is based on the fact that coupled phase oscillators are a basic model
for analyzing synchronization processes in large ensembles of oscillatory sys-
tems: limit-cycle oscillators (see Chap. 6) and chaotic oscillators (see Chap. 7).
Phase dynamics approximation can be successfully applied to any weakly cou-
pled oscillators [7, 8, 20, 186, 187]. Moreover, coupled phase systems are used
to model various systems in physics, engineering, and biology. Among them is
the discrete driven sine-Gordon equation in the under- or overdamped limits,
where the coupling between partial elements has a sinusoidal form. This has
been used as (1) models of coupled pendulum systems [2, 131, 188–193], (2)
Josephson-junction arrays [194–198], (3) magnetic Heisenberg models [199],
(4) one-dimensional (1D) chiral XY model [201], (5) granular superconduc-
tors [200], (6) phase-locked loops [4], (7) phase antenna arrays [202–204], etc.
Thus, their study has also basic importance for applications.
In the first part of this chapter we analyze the collective dynamics in
ensembles of first-order phase oscillators. In the beginning (Sect. 5.1), the
main model of phase dynamics is introduced in a rather general form. Then,
we study collective phenomena in chains of unidirectionally coupled systems
(Sect. 5.2). Section 5.3 is devoted to synchronous regimes in chains with lin-
early and randomly distributed individual frequencies of the rotation. Effects
of nonuniformity of rotations are considered in Sect. 5.4. The classical problem
of mutual synchronization in ensembles of globally coupled phase oscillators is
presented in Sect. 5.5. The second part of this chapter (Sect. 5.6) is devoted to
the study of synchronization phenomena in ensembles of second-order phase
oscillators, i.e., pendulum-like systems. In Sect. 5.7 we conclude this chapter.
104 5 Ensembles of Phase Oscillators
Then, there is an ε0 > 0 such that for all 0 ≤ ε ≤ ε0 the vector of the
phase deviation φ = (φ1 , . . . , φn )T , where T is an operator of transpose, is a
solution to
φ̇j = Hj (φ − φj ) + O(ε), (5.4)
with the vector φ − φj = (φ1 − φj , ..., φn − φj )T , and the function
T0
1
Hj (φ − φj ) = Qj (t)T Gj (γ(t + φ − φj ))dt, (5.5)
T0 0
where Qj (t) is the unique nontrivial T0 -periodic solution to the linear system
ε1 ε1 ε1 ε1
J −1 J J+1
ε2 ε2 ε2 ε2
Fig. 5.1. Schematic coupling of the j th element in the chain. ε1 corresponds to the
coupling strength with the left neighbor while ε2 corresponds to that with the right
neighbor
γ − xj + d1 xj−1 = 0, x0 = 0 (5.17)
leading to ⎧ j
⎨ γ(1 − d1 ) ,
⎪ for d1 = 1,
1 − d1
xj = (5.18)
⎪
⎩
γj, for d1 = 1
Figure 5.2 illustrates the distributions of θj as a function of j for fixed
γ = 0.1 and different d1 (the points for different j are connected by straight
lines).
Note that as follows from (5.16) and (5.18) if γ < 1 and |d1 | < 1, for all
equations in (5.15), there are stable steady states and, therefore, all elements
in the chain of arbitrary size are synchronized (Fig. 5.2b–e).
When γ < 1 and |d1 | ≥ 1 , only some chain elements with 1 ≤ j ≤ j ∗
are synchronized (Fig. 5.2a, f). For the elements with j > j ∗ , nonsynchronous
regimes are generated in the chain.
(2) Let γ > 1 and consider negative coupling strength d1 < 0. Then in the
first two elements of the chain a nonsynchronous regime with a growing phase
difference θ1 develops. What happens for the other phase differences θj (t)?
Depending on the value of the coupling parameter d1 , the rotations along the
chain evolve in different ways. For d1 close to zero and γ slightly larger than
unity, it is possible to suppress the nonsynchronous regime as j grows. In this
case, the nonsynchronous rotatory regime among the first two elements turns
into a nonstrictly synchronous one between the second and third elements,
i.e., θ2 is not constant but oscillates around a certain mean value. Analogous
oscillatory regimes appear between the next subsequent elements, and the
108 5 Ensembles of Phase Oscillators
1.5
(a)
θj
0.0
1.5
(b)
θj
0.0
0.2
(c)
θj
0.1
0.10
(d)
θj
0.05
0.1
(e)
θj
0.0
1.5
(f)
θj
0.0
1.5
0 20 40 60 80 100
j
Fig. 5.2. Stationary phase differences θj for synchronous regimes for γ = 0.1,
N = 100 and different coupling strengths d1 in the model (5.15). (a) d1 = 1, (b)
d1 = 0.9, (c) d1 = 0.5, (d) d1 = −0.5, (e) d1 = −0.9, and (f) d1 = −1.1. Global
synchronization regime sets for (b–e), while cluster synchronization regime (only
for the elements in the beginning of the chain) sets in the cases (a) and (f). In the
latter case, steady states in (5.15) exist only for a few elements in the beginning of
the chain
1.3
(e)
φ20
1.2
1.3
(d)
φ10
1.2
1.3
(c)
φ3
1.2
1.6
(b)
φ2
1.4
1.2
6
(a)
4
φ1
2
0
20000 20020 20040 20060 20080 20100
time
Fig. 5.3. Transition from a nonsynchronous regime to a synchronous one along a
chain (5.15). Time series of phase differences θj for (a) first element, (b) second
element, (c) third element, (d) tenth element, and (e) 20th element. Parameters
are: γ = 1.05, d1 = −0.1
0
0 2π 0 2π 0 2π
θN-1 θN-1 θN-1 θN-1
Fig. 5.4. A schematic scenario for the formation of a quasiperiodic motion with an
increase of coupling d1 in the (θN −1 , θN ) plane. (a) There are stable and unstable
periodic motions. (b) There is only one periodic motion. (c) Quasiperiodic motion.
Only several rotations are shown
5.0
logS(ω)
0.0
5.0
10.0
0 3 60 3 60 3 6 0 3 6 0 3 6
ω ω ω ω ω
Fig. 5.5. Development of instabilities along the chain (5.15). Spectra of the attrac-
tors for several elements. (a) First element, (b) second element, (c) third element,
(d) tenth element, and (e) 20th element. The parameters are γ = 2, d1 = −5, N = 20
which for the phase difference θj = φj+1 − φj and the frequency mismatch
Δj = ωj+1 − ωj yields:
φ̇1 = ω1 + dh(θ1 ), (5.20)
θ̇j = Δj + dh(θj−1 ) + dh(θj+1 ) − 2dh(θj ), (5.21)
where j = 1, ..., N − 1.
5.3 Synchronization Phenomena 113
The conditions for the existence and the stability of the regime of syn-
chronization of all elements in this ensemble, i.e., a regime of global synchro-
nization, can be studied analytically. This regime corresponds to a stable
stationary state in (5.21), which can be defined from the linear difference
equation with constant coefficients:
Δj
xj−1 − 2xj + xj+1 = − , j = 1, ..., N − 1 (5.22)
d
where xj = h(θ̄j ) , θ̄j is the constant difference between the phases of neigh-
bors, and with the boundary conditions:
x0 = xN = 0. (5.23)
N
N
φ̇j = ωj . (5.24)
j=1 j=1
1
N
Ωs = ωj . (5.25)
N j=1
That means that the frequency of the global synchronization regime Ωs for
symmetrically coupled phase oscillators is always equal to the mean frequency
of the oscillators in the chain.
With an increase of the frequency mismatch or with a decrease of coupling
for a certain oscillator j ∗ , xj ∗ reaches a maximal or minimal possible value
(hmax or hmin ) and steady states disappear via a saddle-node bifurcation.
As a result only one phase difference θj ∗ becomes unbounded, i.e., the variable
θj ∗ rotates, while the rest remains bounded, i.e., they oscillate around the
corresponding states θ̄j , i.e.,
Thus, the whole chain is divided into two groups, called clusters, of
mutually synchronized elements. The first cluster consists of the elements
114 5 Ensembles of Phase Oscillators
In the next subsections we use the coupling function h in the form of sine.
Then the main model (5.19) is:
φ̇j = ωj + d sin(φj−1 − φj ) + d sin(φj+1 − φj ) , j = 1, ..., N. (5.30)
First, we consider a linear distribution of the individual frequencies ωj (5.13).
Then using the substitutions θj = φj+1 − φj and Δ = ωj+1 − ωj , the model
(5.30) can be rewritten as:
φ̇1 = ω1 + d sin(θ1 ), (5.31)
θ̇j = Δ + d sin(θj−1 ) − 2d sin(θj ) + d sin(θj+1 ), (5.32)
where j = 1, ..., N − 1.
5.3 Synchronization Phenomena 115
The constant value of Δ allows to solve the linear difference equation (5.22)
analytically. In this case it takes the form:
Δ
xj−1 − 2xj + xj+1 = − , j = 1, ..., N − 1, (5.33)
d
where xj = sin(θ̄j ). From (5.33) the distribution of xj is determined by:
Δ
xj = (N j − j 2 ). (5.34)
2d
Therefore system (5.32) has 2N −1 steady states, whose coordinates θ̄j are
equal to
θ̄j = arcsin[Δ(N j − j 2 )/2d] (5.35)
or
θ̄j = π − arcsin[Δ(N j − j 2 )/2d]. (5.36)
But as was mentioned before, only one steady state with
Δ
θ̄j = arcsin[ (N j − j 2 )]
2d
for all j = 1, ..., N is stable. As follows from (5.34), the steady states disappear
when 2d Δ
(N j −j 2 ) reaches unity. This happens at the critical coupling strength
dcr defined as:
⎧
⎪
⎪ ΔN 2
⎨ 8 , for even N,
dcr = . (5.37)
⎪
⎪
⎩ Δ(N − 1)(N + 1) , for odd N
8
If d becomes smaller than dcr , the break-up of the global synchronization
regime happens, and two clusters of mutually synchronized elements appear.
The size of both clusters is N/2.1 The transitions to (from) global synchro-
nization regime from (to) a fully nonsynchronized regime exhibit interesting
features in dependence on the number of elements and the frequency mis-
match Δ: The main phenomenon is the existence of typical synchronization
trees (Fig. 5.6). Due to the symmetry of the dynamical system, the transition
cascade in these trees is symmetric around the mean individual frequency Ωs .
Similar to two coupled elements, two types of transitions between the cluster
structures are to distinguish: First, a hard transition without intermediate
structures occurs from the state with n clusters to the state with n + 1 or
more clusters or from the state with n + 1 clusters to the state with n or
less clusters. Second, a soft transition happens with a smooth transition of
intermediate structures one into the other. For intermediate structures there
1
We note that here and in all the following chapters we will consider ensembles
with even number of elements.
116 5 Ensembles of Phase Oscillators
1.20
1.15
Ωj
1.10
1.05
1.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6
d
Fig. 5.6. Synchronization trees for the chain (5.30) with linearly distributed indi-
vidual frequencies ωj . The parameters are N = 20, ω1 = 1.0, Δ = 0.01. Both
transition types – hard and soft – are presented. The hard transition is more typical
for relatively large coupling, but the soft transition is quite typical for small coupling
exists the regime of multifrequency rotations, when all elements of the chain
(except the edge ones) rotate with different frequencies Ωj . Scaling properties
of the occurring cluster synchronization regime, namely, sizes of clusters, clus-
ter frequencies, etc., will be described in detail in Chap. 6 for weakly coupled
limit-cycle oscillators.
In Fig. 5.8 we plot space–time diagrams of the evolution of sin(φj (t)) (a-d)
for different synchronization regimes. Corresponding mean frequency (Ωj ) dis-
tributions are shown in Fig. 5.9. In all plots in Fig. 5.8 the darker regions
mark higher values of the presented variables. It is clearly visible that with
an increasing coupling, the collective behavior of the elements in the chain
becomes more coherent, i.e., there appear clusters of mutually synchronized
phase oscillators. So for the coupling d = 1 (Fig. 5.8b) there are five clusters
of synchronized elements. For d = 3 (Fig. 5.8c) there are two synchroniza-
tion clusters. At the border of two neighboring clusters there appear phase
difference slips or defects. These are jumps of the phase difference between
neighboring elements of 2π. In Fig. 5.8a–c these events are indicated by a
merging of two white (black) lines. For regular cluster structures in Fig. 5.8b,
c the frequency of the appearance of defects is defined by the mean frequency
mismatch of the neighboring clusters. For d = 4 (Fig. 5.8d) the regime of
global synchronization is reached. Figure 5.8d demonstrates the propagation
of phase waves with a constant phase shift.
5.3 Synchronization Phenomena 117
0.01
0.00
−0.01
−0.02
Lyapunov exponents
−0.03
−0.04
−0.05
−0.06
−0.07
−0.08
−0.09
−0.10
0.0 0.1 0.2 0.3 0.4 0.5
d
Fig. 5.7. Spectrum of the 20 Lyapunov exponents for a chain (5.30) with linearly
distributed frequencies. Parameters are the same as in Fig. 5.6
4000
3000
2000
1000
0
1 j 50 1 j 50 1 j 50 1 j 50
Fig. 5.8. Space–time plots of evolution sin(φj (t)) for different synchronization
regimes in a chain (5.30) with linear frequency distribution. For the coupling d = 0.5
(a) there are some clusters of synchronization for the elements close to the ends of
the chain and strong incoherent motion for the elements in the middle of the chain.
For d = 1 (b) there are five clusters of synchronized elements. For d = 3 (c) there
are two synchronization clusters. For d = 4 (d) regime of global synchronization is
reached. The parameters are N = 50, ω1 = 1.0, Δ = 0.001
1.40
ε=0.5
1.
1.35
2.
4
1.30
1.25
Ωj
1.20
1.15
1.10
1.05
0 10 20 30 40 50
j
1.10
1.05
Ωj
1.00 1,2
d = 0.07
0.75
02 j 0
3,4,5,6,7,8,9
Ωj
0.90
10,11,12,13,14,15,16,17,18
0.92 d = 0.17
0.80 19,20
Ωj 0.18
12,13,15,16,19,20
0.82
02 j 0
0.70
0.00 0.10 0.20 0.30 0.40 0.50
d
Fig. 5.10. Synchronization trees for a chain with randomly distributed individual
frequencies ωj . The parameters are N = 20, ωj randomly distributed in the inter-
val [0.7 : 1.1]. The numbers of mutually synchronized elements are shown; e.g., 1,2
means that the first and second elements are synchronized. In the inserts the distrib-
utions of the mean frequencies for a constant coupling are presented. The top insert
shows a regime of nonlocal synchronization of the elements 12, 13, 15, 16, 19, 20. The
corresponding place in the synchronization tree is marked by an arrow. The bottom
insert shows the transition of the 18th element from one to another synchronization
cluster
Fig. 5.11. Synchronization in a lattice (5.38) with individual frequencies ωi,j ran-
domly distributed in the interval [0.9 : 1.1]. The observed frequency distribution Ωi,j
for different coupling strength: (a) d = 0, (b) d = 0.015, (c) d = 0.025, (d) d = 0.045,
(e) d = 0.055, (f) d = 0.065, (g) d = 0.075, (h) d = 0.085, (i) d = 0.095. In
all plots the darker regions mark higher values of the observed frequencies. The
regions marked by the same color correspond to clusters of mutually synchronized
elements. Detailed inspections of these results show the existence of all three types
of synchronization behavior observed for chains (Fig. 5.10). The x-axis corresponds
to j = 1, . . . , N = 50 and y-axis to i = 1, . . . , M = 50
simulations we take a < ωj for all j, i.e., in all uncoupled systems there exist
rotations. Hence, the frequency of rotations can be defined according to
ω̄j = (ω1 + Δ(j − 1))2 − a2 (5.40)
0.70 0.48
1.02
0.46 ε=0
0.68 0.1
0.17
1.01
0.44 0.27
0.33
0.37
time scales. For uniformly rotating systems (or also, for example, for qua-
siharmonic limit-cycle oscillators) there exist only one time scale while for
nonuniformly rotating systems (or also, e.g., for relaxation type limit-cycle
oscillators) there are several (at least two) time scales, e.g., a slow and a
fast one. But as it is well known (e.g., [79]) the appearance and interaction
of many timescales lead in oscillatory systems to a chaotic behavior. That
is why the typical route to global synchronization regime in ensembles of
nonuniformly rotators is via soft transitions.
2. During the destruction of the global synchronization regime for all a,
two clusters of mutually synchronized elements appear. As our simula-
tion shows, the size j ∗ of the left appearing cluster becomes smaller with
increasing a.
3. The frequency of global synchronization regime increases with the increase
of the parameter a. For a = 0 it is equal to the mean individual frequency
of the rotators, while for a slightly less than unity, it is close to the maximal
individual frequency. The reason for this effect is the following. The phase
evolution in rotators with large a looks as intermittency of relatively long
intervals of weak changes interrupted by relatively short intervals of strong
increase (see Fig. 2.9). The strong change (“firing”) of the phase variable
in the elements close to the right end of the chain provokes an analogous
strong change of the phases in the neighboring element that leads to a
sequential “firing” in all elements in the chain. Here boundary conditions
have to be taken into account.
4. The frequency of the appearing global synchronization regime Ωs with an
increase of the coupling can be well described by:
Ωs = ωN −j ∗ , (5.41)
0
dcr
d
Fig. 5.13. Typical evolution of the order parameter r according to (5.49)
This equation has been intensively investigated (e.g., [224–229]) and, there-
fore, our network approach may have special interest in the mentioned appli-
cations.
The individual dynamics of a single element (see Sect. 2.5.2), especially
the existence of a parameter region (region D2 in Fig. 2.10) in which two
attractors – a steady state and periodic motion – exist, provides the existence
of two quite different regimes of global synchronization in (5.50):
4.0
3.0 (a)
2.0
.
φj
1.0
0.0
−1.0
4.0
3.0 (b)
2.0
.
φj
1.0
0.0
−1.0
4.0
3.0 (c)
2.0
.
φj
1.0
0.0
−1.0
01 02 03 04 05 06 07 08 09 0 100
j
Fig. 5.14. One (a), two (b) and three (c) solitons in a chain of coupled second-order
phase oscillators in (5.50)
5.7 Conclusions
6.1 Objectives
The investigation of different synchronization phenomena is close to the
problem of the relationship between the properties of temporal dynamics of
oscillations and their changes in spatially extended systems and their chain
analogues. This problem is important for both diagnostics [230, 231] and fea-
sibility of different regimes [232–234]. It is important to emphasize that these
systems have some properties which seem to be surprising and counterintuitive
at the first view: Under certain conditions a time-periodic behavior is real-
ized only in the presence of spatial disorder which is generated either by the
system itself [232] or introduced from outside, e.g., by the dispersion of para-
meters of the elements in the chain [233,234]. In particular, a globally chaotic
regime may be realized in a chain of identical nonlinear regular pendulums
with external forcing under certain conditions. But if there is a dispersion
in their parameters, this chaotic regime is replaced by a periodic one [233].
In an analogous mathematical system, but without external periodic forcing,
it models a parallel chain of current-biased Josephson junctions coupled via
inductors. A small dispersion in the currents gives rise to an enhanced mutual
synchronization of oscillations in these junctions [234].
In this chapter we consider synchronization effects in a chain of coupled
regular oscillators with linearly and randomly distributed individual frequen-
cies (see 5.13). A choice of a completely regular distribution in the form of the
frequency variation that is linear along the chain is motivated by the fact that
it occurs in several situations in engineering and nature as presented next.
ω1 ω2 ω3 ωΝ
Fig. 6.1. Free-ends chain of locally coupled oscillators with different individual
frequencies ωj
ρ̇j = (p − ρ2j )ρj + d(ρj+1 cos θj+1 − 2ρj + ρj−1 cos θj−1 ), (6.7)
j = 1, ..., N
ρj+2 ρj ρj+1 ρj−1
θ̇j = Δ̄j + d sin θj+1 − + sin θj + sin θj−1 . (6.8)
ρj+1 ρj+1 ρj ρj
j = 1, ..., N − 1
ρ2
φ̇1 = Δ1 + d sin θ1 . (6.9)
ρ1
Here, θj = φj+1 − φj is the phase difference and Δ̄j = Δj+1 − Δj is the
frequency mismatch. We consider the following free-end boundary conditions:
ρ0 = ρ1 ; ρN +1 = ρN ; φ0 = φ1 ; φN +1 = φN . (6.10)
φj (T ) − φj (0)
Ωj = φ̇j = lim . (6.11)
T →∞ T
The frequencies Ωj can be also calculated as the 2πnj (T )/T ratio, where
nj (T ) is the number of typical features of the time series (e.g., the maxima
exceeding certain values) in the time interval T . Next, typical solutions of this
model will be presented.
A qualitative picture of the spatiotemporal structure of oscillations is
obtained by plotting shadow-graphs of zj (t) on the (j, t)-plane. Chains of
different number of elements at p = 0.5 were investigated in our numerical
experiments.
The simplest solution of (6.7) and (6.8) is a stable equilibrium state (ρ̇j = 0,
θ̇j = 0) which corresponds to the regime of global synchronization in the
chain, i.e., all oscillators of the chain are in the regime of full synchronization.
In this regime the amplitude equations in a zero approximation (coupling is
134 6 Chains of Coupled Limit-Cycle Oscillators
weak) give the same oscillation amplitudes for all elements of the ensemble.
Then one can use only phase equations (see Sect. 5.3.1) and by taking into
account the condition Δ̄j = Δ, the system of equations for the stationary
phase differences θ̄j is rewritten in the form (see Chap. 5)
Δ
sin θ̄j = (N j − j 2 ). (6.15)
2d
It follows from (6.15) that the system (6.13) has 2N −1 steady states (see,
e.g., [4, 242]) and only one of them (for −π/2 < θ̄j < π/2) is stable. As the
frequency mismatch Δ is increased, the condition of synchronization regime
for all elements:
Δ
(N j − j 2 ) < 1, (6.16)
2d
that coincides with the condition for the existence of equilibrium states [4],
is violated first for j = N/2 at even N , i.e., for the middle element of the
chain. Thus, the condition of global synchronization regime in the chain (or
the respective synchronization area) under the suppositions made above for
the values of the parameters of the initial system is given by the inequality
ΔN 2
8d < 1. (6.17)
such that
Δωc = Δ(N − 1)/2, (6.19)
i.e., the frequency Ωs of the global synchronization regime is equal to the mean
individual frequency in the chain. Moreover (see also Chap. 5), this is true for
all other distributions of the individual frequencies in the quasiharmonic case.
For Δ = 8d/N 2 , we have θ̄N/2 = π/2. In this case, the stable and unsta-
ble equilibrium states merge and a rotatory (with infinite growth of phase
differences θN/2 ) periodic motion is born in the phase space of the system
of equations for the phase difference. Because all the elements of the chain
are coupled, the appearance in the middle element of the regime of unlimited
advance of phase differences between this element and its neighbors leads to
6.2 Synchronization Clusters and Multistability at Linear Variation 135
Ωj
Ωj
0 0 0
0 100 0 100 0 100
j j j
Δ Ωj
Δ Ωj
0 0 0
0.3
d=0.00
d=0.55
d=1.20
0.25 d=3.80
0.2
Ωj
0.15
0.1
0.05
0
0 50 100
j
Fig. 6.3. Perfect cluster synchronization. Averaged frequencies Ωj for different
values of coupling d for Δ = 2 × 10−3
average frequency (Figs. 6.2b and 6.3). The values of the frequency for each
cluster (except the edge ones) are close to those obtained by averaging the
individual frequencies over all the elements forming the cluster. In the case
of a linear dependence of the frequency on j in Fig. 6.3, this corresponds to
the intersection of the lines Ω = Ωj = φ̇j and Ω = jΔ exactly in the middle
of the cluster. These cluster structures are periodic in time – the frequency
differences between the clusters in such structures coincide and are equal to
the lowest cluster frequency in terms of the amplitude equations (6.6):
1.6
1.4
ln(d / d 0) 1.2
0.8
ln (Δ/Δ 0)
0.6
0.4
0.2
0
0.4 0.5 0.6 0.7 0.8 0.9 1
ln Δ Nn
Fig. 6.4. Scaling properties of cluster structures. Critical values of the frequency
gradient (times) in the range Δ ≈ (0.5 − 17) × 10−3 for d = 1 and the coupling
coefficient (plus) in the range d ≈ 0.3 − 3.8 for Δ = 2 × 10−3 , at which the n-cluster
structure breaks prior to the transition from the n to the (n + 1) cluster depending
on the size of the middle cluster Nn . The scale is logarithmic to an accuracy of
arbitrarily chosen origin; the straight lines correspond to the dependence (6.20)
1/2
8d
Nn ∼ . (6.23)
Δ
d=0.8 0.9 1.0 1.1 1.2 1.3 1.45 1.6 1.8 1.8
(a) (b)
Fig. 6.5. Space–time diagrams: (a) intensities of |zj |2 and (b) real parts Re(zj )
for Δ = 0.002 and different values of coupling coefficients. The defects, which are
clearly seen as minima (white regions) of the local amplitude, appear regularly at
certain positions on the chain in the case of perfect cluster structures. The spatial
coordinate j = 1, . . . , 100 is plotted on the abscissa axis, and time t ∈ [0, 4000] on
the ordinate axis
decreasing so that the change of the real part of the complex amplitudes zj in
the (j, t)-plane (Fig. 6.5b) represents correctly the phase of zj . It is important
to emphasize that the formation of a defect in the spatiotemporal pattern of
the phase (or Re(zj )), that is visualized as a singularity of the intensity field
of |zj |2 , corresponds to the transition between the clusters.
Since the number of the defects, nD , formed in one period of a perfect
cluster structure is a unity less than the number of clusters n and their repe-
tition rate is T = 2πΩn−1 , the number of defects per unit time ρD is equal to
6.2 Synchronization Clusters and Multistability at Linear Variation 139
0.5
|z51(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z50(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z49(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z48(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
(a)
0.5
|z51(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z50(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z49(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z48(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
(b)
0.5
|z51(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z50(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z49(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
0.5
|z48(t)|2
0.4
0.3
0.2
0.1
0
50000 51000 52000 53000 54000 55000 56000
(c)
Fig. 6.6. Intensity time series for the middle elements of the chain for Δ = 0.002:
(a) d = 1.45, (b) d = 1.65, and (c) d = 1.8
140 6 Chains of Coupled Limit-Cycle Oscillators
nD Δ(N − 1) n − 1
ρD = = . (6.24)
T 2π n+1
Estimates by these formulas are in very good accordance with the data
obtained directly from numerical solutions. In particular, the number of the
defects is equal to 44, 40, and 39 for the case shown in Fig. 6.5a at d=0.8, 1.2,
and 1.8, and to 45, 42, and 38, respectively, when calculated by the formula
(6.24). Note that, when the transitions between the structures with n and n+1
clusters are caused by changes of the coupling coefficient d, the average defect
density changes only slightly at n ≥ 4. At the same time, their relative position
in the (j, t)-plane alters significantly. For example, in Fig. 6.5a it changes from
completely ordered at d = 1.2 to irregular at d = 1.45, and then again to a
regular one but now with a different symmetry at d = 1.8. The time series
undergo corresponding changes too (see Fig. 6.6).
Now, we discuss some consequences of the scalings of the maximal size of
clusters (6.23) and of the number of defects (6.24) for two limiting transitions
to infinitely long chains N → ∞ at a constant interval of oscillator frequencies
Δω = Δ(N −1) = const. In the first of these two transitions, the thermody-
namic one, the coupling coefficient d remains constant dN = D = const. In the
second one, the “continuous,” the coupling coefficient dN = DN 2 , so that the
corresponding second-order difference in (6.1) and (6.6) tends to the second
derivative with respect to the spatial coordinate. As follows from (6.23), the
maximal size of the clusters in the thermodynamic limit changes as
1/2
8DN
Nn ∼ ,
Δω
so that their relative size N̄n /N and the interval of variations of individual
frequencies along the cluster length, N̄n (Δω )/N , tends to zero as the num-
ber of elements is increased. At the same time, the quantities Nn /N and
N̄n (Δω )/N change approximately as [(8N D)/Δω ]1/2 in the continuous
limit. As a result, the regime of global synchronization N1 = N will inevitably
be established as N → ∞. The mean density of defects in the (j, t)-plane, as
is seen from (6.24), will remain constant in either case, of course if we speak
about the range of the parameters in which the number of clusters is much
larger than unity.
Both the picture of synchronization presented above and its description
in a rather general form on the basis of numerical solutions are possible due
to the high degree of symmetry and homogeneity of the problem in a qua-
siharmonic approximation at small frequency gradients and coupling coeffi-
cients. Actually, the meaningful quantity in this approximation is not the
frequency itself but the frequency difference Δj . Consequently, the system
may be regarded to be a homogeneous one at Δj = Δ = const if the edge
effects are neglected. The picture is becoming more complicated, as Δ and d
6.2 Synchronization Clusters and Multistability at Linear Variation 141
are increased; that makes the effects of multistability and the changes of the
amplitudes of oscillations along the chain essential.
For applications it is important to note that the observed cluster synchro-
nization regime is sufficiently stable under the influence of fluctuations [249].
6.2.4 Multistability
(a)
10
5
n
0
0 0.005 0.01 0.015 0.02
(b)
10
5
n
0
0 0.005 0.01 0.015 0.02 0.025 0.03
(c)
10
5
n
0
0 0.01 0.02 0.03 0.04
Δ
Fig. 6.7. The ranges of frequency gradients at which structures with n perfect
clusters of the type shown in Fig. 6.2b are given for coupling coefficients d = 1 (a);
d = 2 (b); d = 5 (c). Multistability is clearly seen for d = 5 there exist the intervals of
Δ where coexisting (i) 4 and 5 ([∼ 0.005; ∼ 0.007]), (ii) 5 and 6 ([∼ 0.008; ∼ 0.0135]),
(iii) 6 and 7 ([∼ 0.0135; ∼ 0.022]) cluster structures
Δ=0.008
(a) Δ=0.0064
(b)
0.7 0.6
Ωj
Ωj
0 0
0 20 40 60 80 100 0 20 40 60 80 100
j j
Δ=0.0075 Δ=0.006
1.2 1
Ωj
Ωj
0.6 0.5
0 0
0 20 40 60 80 100 0 20 40 60 80 100
j j
Δ=0.007 Δ=0.0055
1.2 0.5
Ωj
Ωj
0.6
0 0
0 20 40 60 80 100 0 20 40 60 80 100
j j
Δ=0.0065 Δ=0.005
1 0.6
0.4
Ωj
Ωj
0.5 0.2
0 0
0 20 40 60 80 100 0 20 40 60 80 100
j j
Δ=0.006 Δ=0.0042
1 0.4
Ωj
Ωj
0.5 0.2
0 0
0 20 40 60 80 100 0 20 40 60 80 100
j j
1.6
Δ = 0.003
Δ = 0.006
Δ = 0.007
1.4 Δ = 0.008
Δ = 0.009
Δ = 0.01
Δ = 0.013
1.2
1
Ωj
0.8
0.6
0.4
0.2
0
0 20 40 60 80 100
j
Fig. 6.9. Nonmonotonic sequence of the number of clusters (4,5,4,5,6,5,6 upward)
at monotonic variation of the frequency gradient Δ and identical initial conditions
(d = 5). The corresponding values of the parameters are marked in Fig. 6.7 by
“times”
We find that the same cluster structure was always established in the region
of the parameters of interest with such variations of initial conditions.
Here, the term containing the oscillation amplitude of the (j − 1)th and
(j + 1)th oscillators, at the breaking of the synchronization regime, may be
regarded as a nonresonant external force, and the term −2dzj that depends
on the magnitude of coupling exerts the same effect as additional losses. If the
losses exceed the amplification (2d > p), there remain only forced oscillations
with amplitude ρj = 2d/|iΔ + p − 2d| that decrease as the mismatch Δ is
increased. Therefore, oscillation death can be observed in the jth element. An
essential aspect in the case of chains is that, with the increase of mismatch Δ,
the synchronization conditions are violated not throughout the whole chain
at once but locally. Namely, in the neighborhood of the weakest element of
the chain, i.e., at the site where the regime of global synchronization breaks
earliest, i.e., in the middle of the chain (see Sect. 6.2.1). In this case, for suffi-
ciently large mismatches, i.e., when the influence of the neighbors is no longer
a resonant one, the coupling acts as effective damping, and for 2d > p the
corresponding element becomes unexcited. As Δ is increased, the region of
oscillation death is expanding so fast that there remain only two clusters that
are not fractioned due to the local desynchronization any longer because the
increase of the parameter Δ/d, that usually leads to a breaking of synchroniza-
tion regime, is compensated by the decrease of the size of the clusters [243].
This is illustrated in Fig. 6.10, where two clusters of synchronized elements
at the edges of the chain are separated by an area of unexcited oscillations.
Therefore, the oscillation death as well as the synchronization is the reason of
7
6 (b)
5
4
Ωj
3
2
1
0
0 20 40 60 80 100
j
0.8
0.7 (a)
0.6
0.5
ρj2
0.4
0.3
0.2
0.1
0
0 20 40 60 80 100
j
Fig. 6.10. Oscillation death. (a) Squared mean amplitudes |ρj |2 and (b) averaged
frequencies Ωj in the case of oscillator death in a chain for d = 5 and Δ = 0.06.
There are two clusters each consisting of 20 mutually synchronized elements at the
ends of the chain
6.4 Effects of Nonuniformity of the Frequency Mismatch Gradient 145
Δ Ωj
Δ Ωj
0.1 0.1
0 0
− 0.1 −0.1
0 20 40 60 80 100 0 20 40 60 80 100
t=3000 t=1000
0.2 0.2
Δ Ωj
0.1
Δ Ωj
0.1
0 0
− 0.1 − 0.1
0 20 40 60 80 100 0 20 40 60 80 100
t=2000 t=500
0.2 0.2
Δ Ωj
0.1
Δ Ωj
0 0.1
− 0.1 0
0 20 40 60 80 100 −0.1
0 20 40 60 80 100
t=0 t=0
0.2
0.2
Δ Ωj
0.1
Δ Ωj
0 0.1
− 0.1 0
0 20 40 60 80 100 − 0.1
0 20 40 60 80 100
j j
(a) (b)
60
50
40
30 1
d
20
2
10
3
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Δ ω*
Fig. 6.12. Average critical values of d at which a transition between different cluster
structures occurs. Region 1 corresponds to the regime of global synchronization.
Two and three clusters of mutually synchronized elements exist in regions 2 and 3,
respectively. The averaging was made over 20 sample random individual frequencies.
Δ = 0.002
1.44
ε=0.3
ε=0.5
ε=0.7
1.43 ε=1.0
1.42
Aj
1.41
1.4
1.39
1.38
0 50 100
j
Fig. 6.13. Mean amplitudes Aj = x2j + ẋ2j for different values of ε for d = 5 and
Δ = 0.00025
1.008
1.006
(b)
Ωj
1.004
1.002
1
0 20 40 60 80 100
j
1.5
1.4
(a)
Aj
1.3
1.2
1.1
0 20 40 60 80 100
j
Fig. 6.14. (a) Mean amplitudes Aj = x2j + ẋ2j and (b) averaged frequencies Ωj
at different values of ε and Δ for d = 5. The symbols are: “plus” for ε = 0.02
and Δ = 0.0006; “times” for ε = 0.1 and Δ = 0.0006; “stars” for ε = 0.3 and
Δ = 0.00065; and “squares” for ε = 0.5 and Δ = 0.0007. The values of Δ are slightly
larger than the critical values Δcr at which the regime of global synchronization is
disturbed
6.5 Synchronization in a Chain of van der Pol Oscillators 149
0.0012
0.001
0.0008
(b)
Δ∗
0.0006
0.0004
0.0002
0 0.5 1 1.5 2 2.5 3
ε
3000
2500
2000
(N* )2
1500
(a)
1000
500
0
0 1 2 3
ε
Fig. 6.15. (a) The dependence on ε of the critical value of mismatch Δ∗ starting
from which the regime of global synchronization is not realized; (b) the dependence
on ε of the squared quantity N ∗ = 50 − Nc , where Nc is the length of the smaller of
the two synchronization clusters that are formed after breaking the regime of global
synchronization, d = 5
150 6 Chains of Coupled Limit-Cycle Oscillators
6.6 Conclusions
– Typical features for the onset and existence of regimes of global (all-to-all)
and cluster (partial) synchronization have been explored.
– Two scenarios, soft and hard, of the transitions between the structures con-
sisting of a different number of synchronization clusters have been revealed.
In the first case, a gradual tuning of the spatial distribution of averaged
frequencies is observed, while in the second one, the transition from the
structure of n synchronized clusters to the structure of n+1 clusters occurs
in a stepwise fashion as a consequence of multistability.
– The effect of different types of individual frequency distributions on syn-
chronization in a constant range of frequency variations is investigated. It
is revealed that the characteristics of the synchronization are strengthened
when an irregular distribution of individual frequencies is used.
– The oscillation death effect has been observed. This effect can also lead to
the creation of synchronized cluster structures.
– Possible mechanisms to control synchronization structures based on the
application of nonhomogeneous forces have been proposed.
– The difference of the transitions to global synchronization regime for weak
and strong nonlinearity has been clarified. For a strong nonlinearity in
the chain with a fixed frequency dispersion, global synchronization regime
occurs at smaller coupling than for weak nonlinearity.
Carrying this out yields linearized equations for the Fourier amplitudes ηk
η̇k = [Df (s) − 4d sin2 (πk/N )E]ηk , (7.4)
where k = 0, 1, . . . , N −1. Solving (7.4) for a special function f (s) and a special
solution s(t), we obtain all Lyapunov exponents (LEs) which determine the
stability of the synchronization manifold M . In [267,268] it was found that for
coupled Rössler systems increasing coupling d can lead to destabilizing of the
synchronous state. This phenomenon is called a short-wavelength bifurcation.
In addition to global and cluster synchronization regimes in ensembles of
identical systems, the phenomena of antiphase and in-phase–antiphase syn-
chronization were observed [270–274]. These types of synchronization are
defined by the existence of stable linear transversal invariant manifolds.
Such antiphase synchronization is observed in a system of coupled oscillators
where all corresponding variables of oscillators are equal with opposite sign.
In in-phase–antiphase synchronization, one set of the corresponding variables
is equal, whereas the other one is equal with opposite signs. Pattern forma-
tion and synchronized chaos was also studied in two-dimensional lattices of
identical Rössler oscillators [275].
ẋj = −ωj yj − zj ,
ẏj = ωj xj + ayj + d(yj+1 − 2yj + yj−1 ), (7.5)
żj = 0.4 + (xj − 8.5)zj ,
j = 1, . . . , N.
Here N is the number of the oscillators in the chain and d is the coupling
coefficient. The parameter ωj corresponds to the natural frequency of the
individual oscillator. Like in the previous chapters, we treat two cases (1) a
linear distribution of natural frequencies ωj = ω1 + Δ(j − 1), where Δ is the
frequency mismatch between neighboring elements and (2) a random distrib-
ution of natural frequencies in the range [ω1 , ω1 + Δ(N − 1)].
We assume again free-end boundary conditions:
Because the Rössler system typically has periodic windows as the parameter
ωj is varied, we choose ω1 and Δ in such a way that at least large periodic
windows are avoided.
In order to study phase synchronization effects here, one needs to have
appropriate phase and frequency definition. In Sect. 2.3 we have proposed a
rather general definition of phase which works well for both phase-coherent
and funnel attractors. Correspondingly, we define for each element in the chain
the phase as:
ẏj
φj = arctan . (7.7)
ẋj
As the phase is well defined, one can straightforwardly calculate the phase
difference between neighboring oscillators θij = φi − φj . This enabled to
analyze synchronization phenomena in chains as follows. If the phase dif-
ference does not grow with time, but remains bounded, i.e.,
we have a 1:1 phase locking between the ith and the jth oscillators in the
chain. Note that these oscillators are not necessary the neighbors. A weaker
condition of synchronous motion is the coincidence of the averaged partial
frequencies:
Ωi = Ωj , (7.9)
which can be calculated as shown in Sect. 2.3.
If
Ωj = Ωs , j = 1, . . . , N (7.10)
a global synchronization regime occurs. In (7.10) Ωs is the common mean
frequency for all elements in the chain, i.e., the frequency of global synchro-
nization regime.
154 7 Phase Synchronization in Ensembles of Chaotic Oscillators
φj =
arctan(yj /xj ),
ρj = x2j + yj2 , (7.11)
zj = zj ,
we obtain
d
θ̇j = Δj + (sin θj−1 − 2 sin θj + sin θj+1 ), (7.13)
2
j = 1, . . . , N − 1,
Δ
sin θ̄j = (N j − j 2 ). (7.14)
2d
and the condition for the onset of global CPS:
ΔN 2
8d < 1, (7.15)
Our computer simulations show very good agreement between theoretical and
numerical results. This supports the idea that synchronization of chaotic oscil-
lators with a rather simple topology, especially, phase-coherent attractors, is
very similar to the case of synchronization of periodic oscillators (Chaps. 5 and
6). However, there is the important difference that the phase difference θj now
is not a constant, but chaotically fluctuates driven by the chaotic behavior of
the amplitudes.
1.039
d=0
0.003
1.038 0.006
0.009
1.037
Ωj
1.036
1.035
1.034
0 4 8 12 16 20
j
Fig. 7.1. Soft transition to global synchronization regime in a chain of Rössler
oscillators (7.5). Averaged frequencies Ωj for different values of coupling d. The
parameters are: N = 20, Δ = 0.0002, and ω1 = 1. The frequency of global synchro-
nization regime is less than the mean individual frequency
0.08
d = 0.003
0.006
0.009
0.06
Lyapunov exponents
0.04
0.02
0.00
−0.02
0 10 20 30 40
k
Fig. 7.2. Forty largest LEs for different coupling for the regimes reported in Fig. 7.1
7.3 Phase Synchronization in a Chain with a Linear Distribution 157
0.04
d = 0.03
0.06
0.18
0.02 0.7
Lyapunov exponents
0.00
−0.02
−0.04
0 10 20 30 40
k
Fig. 7.3. Forty largest LEs in a chain of Rössler oscillators (7.5). The parameters
are N = 50, Δ = 0.009
158 7 Phase Synchronization in Ensembles of Chaotic Oscillators
1.5
1.4 d = 0.0
0.18
0.7
1.3
3.0
5.0
Ωj
1.2
1.1
1
0 10 20 30 40 50
j
Fig. 7.4. Hard transition to global synchronization regime in a chain of Rössler
oscillators (7.5). Averaged frequencies Ωj for different values of coupling d. The
parameters are N = 50, Δ = 0.009, and ω1 = 1
with relatively large frequency difference between the groups. As in the case
of coupled periodic oscillators with a relatively strong frequencies mismatch,
the transition to global synchronization regime and transitions between clus-
ter structures occurs through jumps in the observed frequency distributions.
Hence, we have hard transitions. As a rule, with an increase of d, the clusters
of mutually synchronized oscillators appear rather abruptly. With a further
increase of coupling, the width of the clusters grows in parallel, i.e., the num-
ber of clusters decreases, and, finally, only one cluster remains, and global
synchronization regime is formed. The cluster formation is clearly visible in a
space–time diagram (Fig. 7.5). In all panels a gray scale is used with minimal
values being represented by white and maximal by black. The left panel shows
sin(φj ) = yj /Aj (for system (7.11)); hence the white stripes correspond to the
phase ≈ −π/2 and the black stripes to the phase ≈ π/2. The right panel
shows the amplitudes Aj of the oscillators. To characterize the instantaneous
phase difference between neighboring oscillators, we plot in the center panel
the quantity
φj+1 (t) − φj (t)
sj = sin2 ( ) (7.16)
2
which is zero if the phases are equal and one if they differ by π.
This presentation exhibits the defects, which are clearly seen as maxima
(black regions) of sj and minima (white regions) of the local amplitude. They
appear regularly at certain positions in the chain. Hence, the border between
the clusters is sharp (see Fig. 7.5). Obviously, the frequency difference between
7.3 Phase Synchronization in a Chain with a Linear Distribution 159
The difference in the chain dynamics for small and large frequency mis-
matches directly corresponds to the properties of two interacting systems dis-
cussed in Sect. 4.2.1. First let us mention that a larger frequency mismatch
requires a larger coupling for synchronization regime to occur. We have shown
in Sect. 4.2.1 that for small couplings the frequency difference can be arbi-
trary small; therefore, with an increase of coupling a smooth transition to
synchronization regime is observed in a chain. Contrary to this, for large cou-
plings the frequency difference is either zero or finite; therefore, synchronous
clusters are formed with jumps between them.
Oscillation Death
If the coupling between the elements is not very small, the interaction can
lead not only to synchronization, but also to a suppression of oscillations.
This effect, known as oscillation death, is observed both for pairs and chains
of periodic oscillators (see Sects. 4.1.2 and 6.3). This loss of self-excitation
of two chaotic oscillators due to interaction has been discussed in Sect. 4.2.1.
Here we demonstrate that this effect also appears in a chain of coupled chaotic
oscillators.
To explain it, we rewrite the second equation of (7.5) as
Then it becomes clear that the influence of the coupling can be considered as
some additional damping acting on the system. For large enough frequency
mismatch, the force from the neighboring oscillators is not resonant and does
not compensate the increased losses. As a result, if 2d > a the oscillator can
leave off the self-excited regime, and the oscillations decay or “die out.”
This effect can occur locally in chains of chaotic oscillators, and can go
hand in hand with synchronization. This is illustrated in Fig. 7.6, where a
state with two synchronous clusters near the ends of the lattice is separated
by nonoscillating elements. Note that oscillation death in all oscillators is
also possible.
Fig. 7.6. Oscillation death in a chain of Rössler oscillators (7.5). Space–time dia-
grams of the evolution of yj , sj , and Aj . In the middle of the chain (j=23–38) the
oscillations are suppressed due to the interaction, i.e., oscillation death is observed.
The parameters are N = 50, Δ = 0.015, ω1 = 1, and d = 0.75
1.10
1.08
1.06
1.04
1.02
1.10
1.08
1.06
1.04
1.02
1.10
1.08
1.06
1.04
1.02
1.10
1.08
1.06
1.04
1.02
1.10
1.08
Ωj
1.06
1.04
1.02
0 10 20 30 40 50
j
Fig. 7.7. Averaged frequencies Ωj in a chain of Rössler oscillators with ran-
domly distributed natural frequencies ωj in the interval [1, 1.05]. The number of
elements N = 50, ω1 = 1.0. From bottom to top different coupling strengths
d = 0, 0.01, 0.02, 0.05, 0.2. Effect of nonlocal synchronization is clearly seen for
d = 0.05
1.06
d=0
0.1
0.5
1.04
5
10
Ωj / Ω1
1.02
1.00
0.98
0 10 20 30 40 50
j
Fig. 7.9. Ordering in a chain of identical Rössler oscillators with funnel attractors
(a = 0.23). Space–time diagrams of evolution of yj /Aj , sj , and Aj for different
couplings: (a) d = 0.02 and (b) d = 0.05. The values of yj (t) normalized to the
amplitude are depicted in order to make the phase dynamics visible [yj /Aj = sin φj ]
0.5
a = 0.19
0.21
0.4 0.23
0.25
0.27
0.3
s
0.2
0.1
0.0
0.00 0.05 0.10 0.15 0.20
d
Fig. 7.10. Averaged value of the phase difference s = sj vs. coupling d for dif-
ferent values of the parameter a in a chain of identical Rössler systems with funnel
attractors
As we have shown before for coupled Rössler systems, the tendency to get a
more coherent behavior with an increase of coupling is well pronounced. How-
ever, this tendency is not general. In order to measure the degree of synchro-
nization in a chain of interacting oscillators, we define the frequency disorder
as the standard deviation σ(d) = (Ωj − Ωj )2 of all oscillator frequencies Ωj
depending on the coupling d. For phase-coherent Rössler systems we expect
a decrease with d, as shown for 500 coupled Rössler oscillators (6.5) with
periodic boundary conditions and randomly distributed ωj . Then we analyze
analogous a chain of coupled foodweb oscillators modeled by [31, 281]:
x˙j = a (xj − x0 ) − α1 xj yj
y˙j = −bj (yj − y0 ) + α1 xj yj − α2 yj zj + d
N i (yi − yj ) (7.18)
z˙j = −c (zj − z0 ) + α2 yj zj .
166 7 Phase Synchronization in Ensembles of Chaotic Oscillators
Roessler Foodweb
20
30
10
20
0
y
−10 10
−20 0
−20 −10 0 10 20 0 10 20 30
x x
2.0 1.5
N=500 N=500
1.5
1.0
σ (%)
σ (%)
1.0
0.5
0.5
0.0 0.0
0.0 0.1 0.2 0.0 0.1 0.2
d d
is reminiscent to the Rössler system and therefore one might expect similar
synchronization properties in both systems. To explore this in more detail,
we compare the transition to synchronous regime in coupled chains of Rössler
and foodweb systems. Quenched disorder is introduced by taking bj and ωj
for each oscillator from the same statistical distribution. Despite the fact that
both systems have a very similar attractor topology, we find fundamental
differences in their response to the interaction. For the ensemble of Rössler
systems, the onset of synchronization regime is as expected and σ(d) decreases
monotonically with increasing coupling strength, in accordance to the above
theory. In contrast, the ensemble of foodweb models shows a totally different
behavior. Here, with increasing coupling the frequency disorder is first ampli-
fied leading to a maximal decoherence for intermediate levels of coupling. Only
for much larger coupling strength, the frequency disorder is reduced again and
global synchronization regime sets in. This unusual increase of disorder with
coupling strength is called as anomalous phase synchronization [284]. This
effect emerges because the interaction may perturb the oscillators away from
their attractors. This brings the nonisochronicity – amplitude dependence of
frequency – of the oscillation into play. Disorder enlargement occurs if the
nonisochronicity has positive covariance with the natural frequency of the
oscillation. For the foodweb model this is exactly the case, whereas the oscil-
lations in the phase-coherent Rössler system are practically isochronous.
7.7 Conclusions
In this chapter we have analyzed phase synchronization effects in a chain of
diffusively coupled chaotic Rössler oscillators.
The main findings are:
– When the individual attractor is phase coherent, the dynamics of a chain
is similar to that of a chain of regular oscillators (Chaps. 5 and 6). In a
nonhomogeneous chain, synchronization regime appears when the coupling
exceeds some threshold.
– We have again found two scenarios of synchronization transitions: in the
soft one a gradual adjustment of the mean frequencies is observed, while in
the hard one intermediate clustered states occur. The borders of clusters
appear in the space–time diagrams as positions where phase defects take
place. We have demonstrated that these defects can be both periodic and
irregular. These two scenarios directly correspond to the synchronization
properties of two interacting systems (see Chap. 4): for small coupling the
mean frequencies are adjusted gradually, while for large couplings a virtual
jump is observed.
– In the soft transition global synchronization regime is accompanied by a
high chaoticity of the behavior: the number of the positive LEs is equal to
the number of coupled elements, while in the hard transition the disorder
is strongly reduced: only a few LEs remain positive.
168 7 Phase Synchronization in Ensembles of Chaotic Oscillators
A work similar to the study presented above was done in [276, 277], where
one- and two-dimensional lattices of identical Rössler oscillators have been
considered. Observed in [276], the effect of the appearance of a macroscopic
mean field for very small couplings can be interpreted in our terms as the
appearance of a phase-synchronous state. In the case of nonidentical oscillators
the transition is, however, nontrivial, as we have shown in this chapter.
To conclude, the presented phase synchronization effects in chains of cou-
pled chaotic elements support the idea that phase synchronization is a univer-
sal phenomenon of coupled chaotic systems and is similar to synchronization
in networks of periodic oscillators.
8
Synchronization of Intermittent-Like
Oscillations in Chains of Coupled Maps
where ε is the bifurcation parameter and εcr the critical value for chaos onset.
For coupled maps studied below, the CTS Tc can be calculated numerically as:
1
M
Tc = lim (kl+1 − kl ), (8.2)
M →∞ M
l=1
where kl is the moment when the lth laminar stage sets in or in other words
when the lth firing occurs. We note, that in the studied maps because of
τ /T 1 the time of a full cycle is Tc = τ + T , i.e., the time between the
beginning of two sequential laminar stages is practically equal to τ . Therefore,
the coincidence of averaged τ in such a chain leads to the coincidence of
averaged Tc . One can also introduce a phase of the intermittency, attributing
to each interval between the starts of the laminar stage (or in other words
between two firings) a 2π phase increase:
8.2 Linearly Distributed Control Parameters 171
k − kl
φk = 2π + 2πl, kl ≤ k < kl+1 , (8.3)
kl+1 − kl
where k is discrete time.
The presence of a CTS and a suitable phase allows to formulate the pro-
blem of chaotic phase synchronization in ensembles of coupled units with
intermittent behavior. So, if < τj > or the corresponding frequencies
of all units become equal, this manifests their global 1:1 frequency entrain-
ment. If the conditions
|φkl − φkm | < Const (8.5)
are fulfilled for all k, one can speak about a 1:1 phase locking between the lth
and the mth units.
We demonstrate mutual phase synchronization of chaotic intermittent
oscillations for a chain of diffusively locally coupled nonidentical quadratic
1D maps:
xk+1
j = fj (xkj ) + d(xkj−1 − 2xkj + xkj+1 ),
(8.6)
j = 1, . . . , N,
0.0025
(a) (b) (c)
0.002
0.0015
Ωj
0.001
0.0005
0
0 0.01 0.02 0.03 0 0.01 0.02 0.03 0 0.01 0.02 0.03
d d d
Fig. 8.1. The dependence of Ωj (8.4) on the coupling for ε = 0.000001 and
for three different values of Δε in a chain of 50 coupled maps (8.6) and (8.7).
(a) Δε = 0.000001; (b) Δε = 0.000005; (c) Δε = 0.00001
Fig. 8.1. In all diagrams with an increase of the coupling the tendency for the
formation of a more coherent behavior is clearly seen. Then the parameter
mismatch Δε controls whether a global synchronization regime is created
(Fig. 8.1a) or not (Fig. 8.1b, c). But in all cases the increase of coupling ends
up in a fully incoherent behavior, i.e., a nonsynchronous state. The detailed
analysis of the frequency distribution Ωj vs. coupling (see Fig. 8.2) shows
that the transition to global synchronization regime is smooth, i.e., a gradual
adjustment of frequencies is observed. The reason of such a soft route to a
global synchronization regime is the existence of two quite different timescales:
the slow laminar stage and the fast firing stage. It is well known (see, e.g., [79])
that the appearance and interaction of several timescales (at least two) can
lead in oscillatory systems to a chaotic behavior. Another consequence of the
mixed slow–fast motion is a large value of the frequency in the regime of global
synchronization. It is close to the maximal individual frequency [301]. The rea-
son for this effect is the following: For a sufficiently large coupling, the strong
change (firing) of the dynamical variable in the elements close to the right
end of the chain is faster than in the other elements. This provokes analogous
strong change of the dynamical variable in the neighboring element which also
provokes his neighbor and so on. This process leads to a sequential firing in
all elements in the chain.
8.3 Randomly Distributed Control Parameter 173
0.0008 0.0009
(a) (b)
0.0006
0.0008
Ωj
0.0004
d=0
0.0005 0.0007
0.001
0.0002 0.0015
d=0.01
0.002
0.02
0.0025
0.022
0.0035
0.023
0.0000 0.0006
0 10 20 30 40 50 0 10 20 30 40 50
j j
Fig. 8.2. The dependence of the observed frequencies Ωj on j in (8.6) and
(8.7) for different couplings d for (a) transition “nonsynchronous state–synchronous
state” and (b) transition “nonsynchronous state–synchronous state.” ε = 0.000001,
Δε = 0.0000001, and N = 50
0,0014
(e)
Ωj
0,0013
0,0012
0,00135
(d)
Ωj
0,0013
0,00125
0,0014
(c)
Ωj
0,0012
0,0014 (b)
Ωj
0,0012
0,0015
(a)
Ωj
0,001
0 10 20 30 40 50
j
Fig. 8.3. The dependence of the observed frequencies Ωj on j in (8.6) and (8.7) for
different couplings (a) d = 0, (b) d = 0.0005, (c) d = 0.001, (d) d = 0.0015, and (e)
d = 0.0025. ε = 0.000001, Δε = 0.0000001, and N = 50
15000
(a) (b) (c) (d) (e)
1
1 j 50 1 j 50 1 j 50 1 j 50 1 j 50
of fully developed STI sets in. This rich spatiotemporal dynamics in the syn-
chronous and nonsynchronous regimes is illustrated in Fig. 8.4. The left panel
corresponds to a nonsynchronous behavior (small values of coupling), but
there are already several clusters of mutually synchronized elements. Only
176 8 Synchronization of Intermittent-Like Oscillations
panel (b) corresponds to a synchronous regime. In all plots the darker regions
mark higher values of the presented variables.
Next, we analyze these processes observed by using our phase definition
(8.3). This way the regimes of perfect (Fig. 8.4b) and imperfect (Fig. 8.4c)
chaotic phase synchronization are characterized by a phase distribution φj ,
which is a sequence of intervals with constant phase, separated by ±2π-kinks.
The position of the kinks at constant time corresponds to phase slips. In the
synchronous regimes the phase slips appear with the frequency of synchro-
nous motion. In the nonsynchronous regimes phase slips appear randomly
and rather fast.
In the presented model STI appears due to the relatively strong interaction
of many units. The specific property in our observation consists in the exis-
tence of a transition from a fully coherent (synchronous laminar) to a fully
noncoherent (nonsynchronous turbulent) behavior. In order to demonstrate
this transition, we plot in Fig. 8.5 the ratio D of the number of laminar stages
corresponding to the synchronization regime and the full number of laminar
stages. It is clearly seen that (1) for d ≥ d∗cr the turbulent stages appear very
rarely, whereas (2) for d ≤ d∗∗
cr there are very short intervals of laminar stages.
In our numerical study we also examined chains of different sizes and
different boundary conditions and find that qualitatively all described above
effects are the same.
0,8
0,6
D
0,4
0,2
0
0,05 0,06 0,07
d
Fig. 8.5. The dependence of the ratio D on the coupling for a chain of 50 elements
(8.6) and (8.7) with εj randomly distributed in the interval [0.000005; 0.000015].
d∗cr ≈ 0.049 and d∗∗
cr ≈ 0.067
8.4 Collective Oscillations in a Chain of Spiking Maps 177
where xj and yj are the fast and slow variables, respectively. μ = 10−3 and σj
are the parameters of the individual map and d is the coupling. The function
f (·, ·, ·) has the form:
⎧
⎪
⎪ α/(1 − xk ) + y k , if xk ≤ 0,
⎪
⎪
⎪
⎪
⎪
⎪
⎨ α + y k , if 0 < xk <
k
f (x , x , y ) = α + y k and xk−1 ≤ 0,
k−1 k
(8.10)
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪ −1, if xk ≥ α + y k
⎩
or xk−1 > 0
Fig. 8.6. Space–time plots of xj in (8.9) and (8.10) for synchronous (b) and non-
synchronous regimes (a, c, d) for σj randomly distributed in the interval [0.15; 0.16].
N = 100, d = 0.005 (a), d = 0.05 (b), d = 0.09 (c), d = 0.2 (d) and (e). Panel (e)
is an enlargement of panel (d)
bursts (Fig. 8.6d, e). Note, that spikes forming these bursts are correlated in
space, as they appear as triangular embedding with a fractal-like spatiotempo-
ral structure. The transition observed shows how spiking maps can produce
bursting behavior if they form a spatially extended system. Why collective
complex behavior differs for intermittent and spiking maps? This is due to
the interplay between fast and slow dynamics that produces spiking behavior
in system (8.9) and (8.10). The slow variable regulates the threshold value and
when the threshold gets too high, it forces spike events to stop propagating
along the chain and the burst ends. Until the fastest neuron is recovered, no
spiking is observed in the chain and that separates bursts clearly. Quite on the
contrary, there is no slow variable in the intermittent map that would regulate
turbulent outbursts and they multiply freely in the regime of spatiotemporal
chaos.
First we show that the transition to mutual phase synchronization takes place
on the bursting timescale of the globally coupled oscillators, while on the
spiking timescale they behave asynchronously. The onset of mutually syn-
chronized bursting in the studied ensemble has much in common with the
classical example of global phase entrainment of phase oscillators (Chap. 5).
The transition to synchronized bursting is observed as the coupling between
the oscillators is increased (Fig. 8.7b). A nonzero mean field is also formed and
its oscillations make neurons develop a common rhythm. Remarkably, only the
slow timescale (i.e., bursting) dynamics becomes coherent. The spikes remain
uncorrelated and do not substantially contribute to the mean-field (close to
periodic) dynamics. The frequencies Ωi of the bursting oscillators calculated
for different coupling illustrate the appearance of one synchronized cluster and
its gradual increase in size (Fig. 8.8). The asymptotic (very large n) behav-
N
ior of the order parameter r = | eiφ(j,n) |/N indicates a second-order phase
j=1
transition to coherence (Fig. 8.9), which is typical for mean-field coupled phase
oscillators [8, 305] (see Sect. 5.5).
mean field
mean field
α=4.4 α=4.4
α=4.1 α=4.1
(a) (b)
0 1000 2000 0 1000 2000
n n
Fig. 8.7. Time series x(j, n) of two neurons (8.11) and the mean field from
(a) an uncoupled ensemble (ε = 0) and (b) a coupled ensemble (ε = 0.04), where
synchronization regime of the bursts is achieved, in the absence of an external signal
(A = 0). Random values of αj are implemented, σ = β = 0.001, N =1,000
8.5 Synchronization in Ensembles of Globally Coupled Bursting Oscillators 181
0.03
ε=0.025
Ωj 0.02
0.01
0.016 0.018 0.02 0.022 0.024 0.026 0.028 0.03 0.032
0.024
ε=0.035
Ωj
0.022
0.02
0.016 0.018 0.02 0.022 0.024 0.026 0.028 0.03 0.032
0.0215
ε=0.045
Ωj
0.021
0.016 0.018 0.02 0.022 0.024 0.026 0.028 0.03 0.032
Ω0j
Fig. 8.8. Frequencies of bursting in the mean-field coupled ensemble (8.11) vs. those
at the zero mean-field coupling show a growth of the synchronization cluster as the
coupling coefficient ε is gradually increased. The external signal is absent (A = 0),
N =1,000
0.9
0.8
0.7
0.6
0.5
r
0.4
0.3
0.2
0.1
0
0 0.01 0.02 0.03 0.04 0.05 0.06
ε
Fig. 8.9. The order parameter r vs. the mean-field coupling coefficient ε indicates
a second-order phase transition to CPS of bursting in (8.11), N =1,000
182 8 Synchronization of Intermittent-Like Oscillations
ε
N
∗ αj ∗ ∗
x(j , n + 1) = + y(j , n) + x(i, n) + A sin ωn (8.12)
1 + x(j ∗ , n)2 N i=1
The equations for the y variable of this neuron and those for the other neu-
rons remain unchanged. The number of bursting oscillators in the simulated
ensemble is N = 50. By applying a frequency of the driving signal ω which
differs a bit from those of the mean field, we obtain external phase locking
(i.e., external CPS) of bursts (Fig. 8.10) in dependence on the forcing ampli-
tude A. This transition to external CPS in this ensemble is characterized by
the following main properties:
x 10−3
2.5
2 A=0.05
A=0.09
1.5
A=0.15
1
0.5
Ωj − ω
−0.5
−1
−1.5
−2
−2.5
0.014 0.015 0.016 0.017 0.018
ω
Fig. 8.10. The difference between frequencies of bursting in oscillators and the
driving frequency vs. the driving frequency for three values of the driving amplitude
demonstrates external CPS in the ensemble of bursting oscillators (8.11) and (8.12).
The fixed mean-field coupling ε = 0.1 ensures mutual phase synchronization regime
between oscillators in the absence of the driving signal, N = 50
8.5 Synchronization in Ensembles of Globally Coupled Bursting Oscillators 183
450
400
350 ω=0.0158
300
250
nω − φj
200 ω=0.0156
150
100
50 ω=0.0153
0
−50
0 2 4 6 8 10
n x 105
Fig. 8.11. Drifting (ω = 0.0158 and ω = 0.0156) and locked (ω = 0.0153) phases
φj bursting oscillators illustrate the transition to CPS in a neural ensemble (8.11)
and (8.12). For each frequency ω the differences nω − φj for all j = 1, . . . , N are
presented. Some lines practically coincide, N = 50
184 8 Synchronization of Intermittent-Like Oscillations
quiescent regime, when positive, and delay it, when negative, which tends to
synchronize the driven neuron. The mean-field coupling term reflects averaged
individual dynamics of the neurons. Suppose, that the external frequency
exceeds that of the mutually synchronized autonomous ensemble. Then the
periodic signal will fasten the oscillations of the driven neuron. When the
whole ensemble is about to start (stop) bursting, the global dynamics becomes
very sensitive to changes of the amplitudes of the individual oscillators. If the
fastened neuron starts (stops) bursting (passing ahead of the others), the
abrupt change of its amplitude increases (decreases) the mean-field value,
pushing the other neurons toward bursting (silence). Thus higher frequencies
win the competition with the lower ones in the neural ensemble. Quite on the
opposite, should the frequency of the driving signal be smaller than that of
the autonomous ensemble, only tiny synchronization effects can be expected.
From this the frequency asymmetry of the synchronization plateaus results.
We would like to stress that a local driving can result in external CPS of the
whole ensemble only when the oscillations in the autonomous ensemble are
mutually synchronized. Oscillators that are not bursting coherently with the
driven oscillator are not susceptible to the driving signal. The contribution of
the driven oscillator to the mean field is proportional to ε/N and does not
depend upon the amplitude of the driving A. That explains why the increase
of the synchronization region is limited when A is increased and ε is fixed. It
x 10−3
1.2
1
Ndr =4
0.8
Ndr =2
Δω
0.6
0.4
Ndr =1
0.2
0
0.005 0.01 0.015 0.02
1/N
Fig. 8.12. The average size of the synchronization plateau Δω (see details in the
text) for the fixed mean-field coupling ε = 0.2 and the driving amplitude A = 0.15
vs. the inverse number of oscillators in the ensemble 1/N for different numbers of
driving neurons Ndr
8.6 Conclusions 185
also follows that the synchronization region should decrease as the number of
oscillators in the ensemble grows. To analyze this dependence, we calculate
the size of the synchronization plateau Δω for a fixed mean-field coupling
ε = 0.2 and driving amplitude A = 0.15 in ensembles with a different number
of oscillators N . For each size value N we generate 100 realizations of random
αj , j = 1, . . . , N and average the obtained sizes of synchronization regions. In
Fig. 8.12 we observe that the synchronization region size scales as: Δω ∝ 1/N .
For control purposes in very large ensembles one can apply the same driving
signal not to one but to several arbitrary taken neurons Ndr (see Fig. 8.12 for
two and four driving neurons). The increase of the number of drivers leads to
the increase of synchronization plateaus.
8.6 Conclusions
In this chapter we have analyzed phase synchronization phenomena in chains
of coupled chaotic intermittent maps.
The main properties are:
In this chapter we study conditions for an onset of regular and chaotic phase
synchronization (PS) regimes in ensembles of coupled circle maps (CMs) [183].
For networks of coupled maps different problems of synchronization, pat-
tern formation, and spatiotemporal chaos have been investigated [306–311].
In most studies, however, identical coupled maps have been analyzed. It
is evident that for networks of coupled identical elements the investigation
of PS makes no sense, because the individual frequencies of the uncoupled
elements already coincide. Therefore we consider here mainly coupled non-
identical maps, i.e., the more realistic case that usually arises in nature and
engineering where subsystems are never identical. In contrast to other maps,
CMs have the strong advantage that the phase variables exist which allows
immediately to apply criteria to test for of synchronization similar to those
used for the detection of PS in time-continuous systems (see Chaps. 5–7).
As for other phase systems, synchronization in ensembles of coupled CMs
has found important practical applications in electronics, radioengineering
or communications, in particular, in networks of digital phase-locked loops
(DPLLs) [4, 80, 81, 312, 313]. Ensembles of coupled CMs can be used as rather
simple but paradigmatic models to investigate processes of mutual synchro-
nization in ensembles of relaxation time-continuous dynamical systems (see
Chap. 6). In this case, each value of the phase variable can be interpreted as
an onset of a new impulse (or firing event) [1, 314, 315]. The systems treated
in this chapter belong to the broad class of “pulse-coupled” systems arising
in many branches of science and engineering; e.g., pulse-coupled systems have
been investigated as models of neural networks [187, 316–323], cardiac pace-
maker cells [324–326], or in communication [327].
The chapter is organized as follows. In Sect. 9.1 we present a model of
chains of coupled CMs. Then we consider synchronization effects in ensembles
of identical circle maps (Sect. 9.2). This will be helpful for the study of collec-
tive behavior in nonhomogeneous chains that has been done in the rest of this
chapter. In Sect. 9.3 we state the problem and discuss criteria of synchronous
188 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
behavior. Sections 9.4 and 9.5 are devoted to the synchronization of regular
and chaotic CMs, respectively. The results are summarized in Sect. 9.6.
φk+1
j = ωj + φkj − F (φkj ) +
(9.1)
d1 sin (φkj−1 − φkj ) + d2 sin (φkj+1 − φkj ).
φkj
F (φkj ) = c (9.2)
π
defined in the interval (−π, π] and c is the control parameter. In the following
we use for the ensembles basic properties of a single circle map (Sect. 2.5.4)
and synchronization effects for two coupled CMs (Sect. 4.3).
The parameters ωj characterize the individual frequencies. For simplicity,
we call ωj as frequency.
We assume that the system is subjected to:
Let us briefly discuss our choice of coupling. System (9.1) with the
nonlinear coupling can be regarded as a model of a multichannel chain of
partial DPLLs connected in parallel by phase-mismatching signals [329]. To
realize these connections in a chain in its simplest variant, it is necessary
to compare the output signals of two neighboring DPLLs with the help of a
separate phase discriminator (PDs) and then to apply the obtained phase-
mismatching signal for the frequency control of both generators (Fig. 9.1).
Some similar one- and two-dimensional space models of coupled identical CMs
have been studied in [4].
9.2 Synchronization in a Chain of Identical Circle Maps 189
PDj−1 PDj
Fig. 9.1. Digital phase-locked loops (DPPLs) coupled via phase discriminators
(PDs)
We analyze the nonlinear coupling between the partial elements in the form
of sine of the phase differences because this kind of coupling naturally arises
in models of ensembles of weakly locally diffusively coupled time-continuous
oscillators (see Chap. 5). In contrast to often used types of diffusive coupling
like linear phase difference between the neighbors:
or through the same nonlinear functions as individual functions for each ele-
ment:
d1 (F [φkj−1 ] − F [φkj ]) + d2 (F [φkj+1 ] − F [φkj ]), (9.6)
the sine type coupling exhibits some special properties of the dynamics of
populations of time discrete elements. The most important advantage of the
sine coupling (9.1) is that it generates mutually phase synchronous rotations
already for a very small coupling d compared to the cases (9.5) and (9.6).
In this section we study the case of identical maps, i.e., ωj = ω, in the chaotic
regime, i.e., c < 0. In homogeneous chains of chaotic oscillators the regime
of complete chaotic synchronization is possible. This regime is realized if the
state
φk1 = φk2 = . . . = φkN = φ̄k (9.7)
is stable.
The specific type of the considered function F (φ) (9.2) allows to find
analytically the stability conditions of the synchronous state.
We will consider symmetrically and asymmetrically coupled maps.
190 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
ρl = ρ0 − 2d(1 − cos ψl )
(9.10)
l = 0, 1, . . . , N − 1,
where ψl = 2πl/N . If d = 0, i.e., the maps are uncoupled, hence all N eigen-
values are equal to ρ0 . With increasing d, all eigenvalues except ρ0 become
smaller and go into the interval (−1; 1). The largest eigenvalue among them
are ρ1 = ρN −1 . The smallest eigenvalues are ρN/2 for even N and ρ(N ±1)/2 for
odd N . Therefore, in order to obtain the stability region of the synchronous
state, it is sufficient to look on the eigenvalues ρ1 and ρN/2 for even N , and
ρ1 and ρ(N +1)/2 for odd N . Then the synchronization region in (9.1) can be
determined from the inequalities:
9.2 Synchronization in a Chain of Identical Circle Maps 191
1. For even N
1
d∗∗ ≡ c
2π(1−cos(2π/N )) <d< 2 − c
4π ≡ d∗ (9.12)
2. For odd N
2π−c
d∗∗ ≡ c
2π(1−cos(2π/N )) <d< 4π(1−cos(π[N +1]/N )) ≡ d∗ . (9.13)
These conditions show that the stability of the synchronous state is com-
pletely determined by (a) the eigenvalue of the single map and (b) by the
number of coupled maps in the ensemble. In Fig. 9.2 the dependencies of d∗∗
and d∗ on the parameter c for different N is shown. One can see that with
increasing N the synchronization region Sp become smaller.
Assuming a chain of fixed length N , one can calculate the maximal value of
ccr , such that at c ≤ ccr a synchronous regime does not exist for any coupling.
Sp Sf Sf Sf
Sp Sp
0.4 0.4 0.4
d
0 0 0
− 0.8 − 0.6 − 0.4 − 0.2 0 − 0.2 − 0.15 − 0.1 − 0.05 0 − 0.03 − 0.02 − 0.01 0
c c c
Fig. 9.2. Synchronization regions in the chain (9.1) with periodic boundary condi-
tions (region Sp bounded by dashed lines) and free-end boundary conditions (region
Sf bounded by solid lines) for N = 10 (a), N = 20 (b), N = 50 (c). Note different
scales for the horizontal axis
192 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
10
periodic BC
free-ends BC
1
−ccr
0.1
0.01
0.001
0 10 20 30 40 50 60 70 80 90 100
N
Fig. 9.3. Dependence of the critical parameter value −ccr on the number of elements
N in the chain (9.1) with periodic and free-end boundary conditions. Synchroniza-
tion takes place in regions under the curves
This value can be easily found from 9.12 and 9.13. For example, for even N
it reads (Fig. 9.3):
cos(2π/N ) − 1
ccr = 2π (9.14)
cos(2π/N ) + 1
With increasing coupling the last unstable eigenmode to be observed when
approaching synchronization region from below for d a bit less than d∗∗ has
the wave number ψ1 = 2π/N and the spatial period equal to N . Hence, a
long-wave bifurcation takes place. The distribution of φkj for fixed k will be
sinusoidal: a sin(2πj/N ) + b, where a and b depend on k.
For even N the first mode which becomes unstable at the transition from
a synchronous to a nonsynchronous regime is the eigenmode corresponding to
l = N/2, ψ1 = 2π/N , and spatial period equal to 2. Therefore for d a bit larger
than d∗ the “antiphase” (“saw”) distribution occurs, at which the values φk
for all k are identical for all even elements, i.e., φk2j = φ̄keven as well as for all
odd elements, i.e., φk2j−1 = φ̄kodd . Hence, a short-wave bifurcation takes place.
These and other distributions φj in the chain (9.1) with periodic boundary
conditions at N = 256, c = −0.0001 for different couplings d after 10,000,000
iterations are shown in Fig. 9.4. For chosen parameters with increasing d
the transition “nonsynchronous regime–synchronous regime” occurs at d∗∗ ≈
0.2113. The transition “synchronous regime–nonsynchronous regime” takes
place at d∗ ≈ 0.5. Sinusoidal and “antiphase” distributions of φj are presented
9.2 Synchronization in a Chain of Identical Circle Maps 193
φj − 0.8136 (a)
− 0.8137
− 0.8138
− 0.6
(b)
− 0.8
φj
−1
(c)
− 0.06186
φj
− 0.06188
2
(d)
0
φj
−2
−4
2
(e)
0
φj
−2
0 64 128 192 256
j
Fig. 9.4. Distributions φj in the chain (9.1) with periodic boundary conditions
for N = 256, c = −0.0001 at different coupling strength d after 10,000,000 iter-
ations. d = 0.01 (solid line), d = 0.04 (dotted line corresponds to distribution
φj = a sin(2πj/N ) + b, d = 0.07 (dashed line corresponds to homogeneous syn-
chronous state) (a), d = 0.508 (“antiphase” distribution) (b), d = 0.748 (c) (values
φj only for even j are shown), d = 0.88 (d), d = 0.91 (e)
3
d=0.504
d=0.509
2
1
φj
−1
−2
−3
0 20 40 60 80 100
j
Fig. 9.5. Distribution φj in the chain (9.1) with periodic boundary conditions at
N = 100, c = −0.0001 for different couplings d after 10,000,000 iterations. At
d = 0.504 the distribution φj is linear: φj = a + 2πj/N . At d = 0.505 “antiphase”
regime is stable
Such regions Sf for different N are presented in Fig. 9.2 which shows that for
chains of fixed length the stability region of the synchronous state in chains
with periodic boundary conditions Sp is essentially larger compared to those
with free-end boundary conditions Sf .
The critical value ccr is:
1
ccr = π(1 − ) (9.17)
cos(π/N )
This curve lies below the curve ccr (9.14) for periodic boundary conditions
(Fig. 9.3), i.e., we find again that for chains closed in a ring the synchronous
regime can exist for larger size (larger N ) than in chains with free ends.
In Fig. 9.6 the distributions of φj in the chain (9.1) with free-end boundary
conditions for N = 256, c = −0.0001 for different d after 10,000,000 iterations
are shown. At the transition to a synchronous regime at d a bit less than d∗∗ ,
the last stable eigenmode corresponding to l = 1 has the wave number π/N
9.2 Synchronization in a Chain of Identical Circle Maps 195
− 0.8267 (a)
− 0.8268
− 0.8269
(b)
0.590509
0.590508
− 0.76
(c)
− 0.8
− 0.84
− 0.88
2
(d)
0
−2
2
(e)
0
φj
−2
0 64 128 192 256
j
Fig. 9.6. Distributions φj in the chain (9.1) with free ends for N = 256, c = −0.0001
for different coupling strength d after 10,000,000 iterations. d = 0.01 (a), d = 0.211
(b) (distribution φj = a sin(πj/N ) + b, d = 0.501 (c), d = 0.864 (d), d = 0.99 (e)
In the case of asymmetrically coupled maps (9.1) the stability conditions can
be also found analytically.
For periodic boundary conditions the eigenvalues are complex:
and ρ1 and ρ(N ±1)/2 for odd N . Hence, the boundaries of the stability regions
can be obtained from the equalities:
which define the stability region for even N in the coupling parameter plane
(d1 , d2 ) by the curves:
2π
ρ20 − 2(1 − cos )∗
N
2π
∗[d∗∗ ∗∗ ∗∗ 2 ∗∗ 2 ∗∗ ∗∗
1 + d2 + (d1 ) + (d2 ) − 2d1 d2 cos ]=1 (9.20)
N
ρ0 + 1
d∗1 + d∗2 = .
2
For free-end boundary conditions the eigenvalues are:
√
ρl = ρ0 − (d1 + d2 ) + 2 d1 d2 cos ψl
(9.21)
l = 0, 1, . . . , N − 1,
where ψl = πl/N . The stability region of the synchronous state defined from:
ρ1 = 1; ρN −1 = −1 (9.22)
2 2 2
(a) (b) (c)
1.5 1.5 1.5
S
d2
1 1 1
1 1 1
S S
0.5 0.5 0.5
S
S S
0 0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2 0 0.5 1 1.5 2
d1 d1 d1
(9.25)
l1 = 0, 1, . . . , N − 1, l2 = 0, 1, . . . , M − 1
where ρ0,0 is the multiplier of individual map, ψl1 = 2πl1 /N and ψl2 =
2πl2 /M .
Let us focus on the case of even N and M . Then stability region of synchro-
nous regime can be found from the conditions ρ1,1 = 1 and ρN/2,M/2 = −1.
It is bounded by:
Fig. 9.9. Spatial structures in the lattice (9.24) of coupled CMs with periodic
boundary conditions for ω = 0.5, c = −0.001, N = M = 20, and d = 10−6 (a),
d = 4.2 · 10−5 (b), d = 0.28 (c) and d = 0.5 (d)
Theory
φk+1
1 = ω1 + φk1 + d sin θ1k ,
(9.32)
j = 1, . . . , N − 1 (9.33)
with Δ = ωj+1 − ωj , θjk = φkj+1 − φkj and the boundary conditions: θ0k =
k
θN = 0.
The conditions for the onset of a global synchronization regime and the
frequency of globally synchronized rotations can be found analytically in anal-
ogy with the case of time-continuous uniformly rotating phase oscillators (see
Chap. 5). The stable fixed point θjk+1 = θjk = θ̄j for each j = 1, . . . , N − 1 in
system (9.33) corresponds to a regime of global synchronization in the chain.
Then the system of equations for the stationary phase differences θ̄n can be
written as (compare with 5.19 and 6.13–6.15):
As follows from Chap. 5 (see (5.33) and (5.34)), the distribution of θ̄j is
Δ
sin θ̄j = (N j − j 2 ). (9.35)
2d
It follows from (9.35) that the system (9.33) can have 2N −1 fixed points, but
only one of them (θ¯j ∈ [−π/2; π/2] for all j = 1, . . . , N − 1) is stable. As the
frequency mismatch Δ is increased, the condition for the existence of a stable
fixed point is
Δ
| (N j − j 2 )| < 1. (9.36)
2d
202 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
This is violated firstly for j = N/2 at even N , i.e., for the middle element
in the chain. Thus, the condition for the existence of a fixed point in the
N -element chain is again given by the inequality
ΔN 2
| | < 1. (9.37)
8d
This condition for the existence of a global synchronization regime in system
(9.1) coincides with the results of our numerical experiments. The rotation
number of synchronous rotations ρs can be determined from (9.32):
φk+1
1 = ω1 + φk1 + d sin θ̄1k =
(9.38)
Δ(N −1)
ω1 + φk1 + 2
and is equal to the mean rotation number over all elements in the ensemble
Δ(N − 1)
ρs = ω1 + . (9.39)
2
With increasing of the frequency mismatch Δ, a loss of global synchronization
regime takes place. For a long chain a two cluster synchronization occurs, i.e.,
the chain is divided into two clusters of equal sizes that consist of mutually
synchronized CMs of different rotation numbers.
Numerical Results
Fig. 9.10. Transition trees of synchronization in system (9.1) for the rotation num-
bers ρj vs. the coupling d for different frequency mismatches Δ and for uniform
rotations c = 0, ω1 = 0.1. (a) Δ = 0.001, (b) Δ = 0.0001, and (c) Δ = 0.00001,
N = 100
(e.g., 4 to 6) and (2) through a merging of two clusters into one (e.g., 6 to 3).
The values of the rotation number for each cluster (except the edge ones) are
close to those obtained by averaging the individual rotation numbers over all
elements forming the cluster. For a weak coupling this interstructural transi-
tion is soft: A gradual adjustment of the rotation numbers takes place. For
such soft transitions most elements of the chain (except the edge ones) rotate
with different rotation numbers.
It should be noted that the transition n-cluster structure − > nonsynchro-
nized state − > again n-cluster structure is quite typical. Usually at such
transitions the number of elements in the clusters is changed. Transition trees
of synchronization are very similar not only qualitatively but they also, as
found numerically, exhibit well-expressed scaling properties. So, if in the N -
element chain with the frequency mismatch Δ1 some cluster structure appears
(disappears) at the coupling dcr
Δ1 , then the same cluster structure appears (dis-
appears) in the same chain but with another Δ2 at the coupling:
Δ1
dcr cr
Δ 2 = dΔ 1 . (9.40)
Δ2
204 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
Therefore, knowing the evolution of the rotation number distribution for some
Δ1 , one can easily calculate a similar evolution for any other value of frequency
mismatch Δ2 .
As in the case of two coupled CMs, the global synchronization regime can
disappear with the increase of coupling. The stable fixed point of system (9.34)
corresponding to the global synchronization regime at some critical coupling
d, loses its stability through a period doubling. But if all θj which become
oscillating around
Δ
θ̄j = arcsin (N j − j 2 ) (9.41)
2d
remain bounded, global synchronization regime still exists. Figure 9.10 does
not show this change, because there the averaged rotation numbers are plot-
ted. But when coupling becomes larger than some coupling d∗ , the oscillatory
behavior of θj no longer exists and the global synchronization regime disap-
pears. Note that not only the regime of global synchronization can be changed
into the regime of global nonsynchronization. Figure 9.10a presents an exam-
ple of a nonsuccessful transition to global synchronization regime in a chain
with a relatively large frequency mismatch Δ. The two-cluster structure is
immediately transformed into fully incoherent rotations.
ρj
0.097 0.097 0.095 0.095
ρj
0.097 0.097 0.095 0.095
ρj
0.096 0.096 0.095 0.095
ρj
0.096 0.096 0.094 0.094
ρj
0.096 0.096 0.093 0.093
ρj
0.095 0.095 0.093 0.093
0 25 50 0 25 50 0 25 50 0 25 50
i i i i
elements can have the same rotation numbers, i.e., they belong to one interval
of equal rotation numbers (see Fig. 2.12). Thus, there are clusters for which
the rotation numbers coincidence of rotation numbers without coupling. If the
difference between the rotation numbers of the elements is small enough, the
occurrence of a global synchronization regime happens for smaller coupling
than in the case of weak noncoherence. There, the common rotation num-
ber coincides with the rotation number of the elements in the largest cluster.
Usually this situation is observed for intermediate values of c. For strong non-
coherence, the rotation number difference can become very large. Then in spite
of the existence of clusters with coincident ρj in an uncoupled chain, global
synchronization regime can be observed only for stronger coupling (Fig. 9.12).
206 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
0.8
0.7
0.6
0.5
d+
0.4 b1=1.0
0.3 1.2
0.2 1.6
1.8
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4
c
N
N
(φk+1
j − φkj ) = ωj . (9.42)
j=1 j=1
Because in the global synchronization regime the rotation numbers of all ele-
ments coincide, i.e., ρ1 = . . . = ρj = . . . = ρN = ρs , (9.42) yields
1
N
ρs = ωj . (9.43)
N j=1
This means that the rotation number of the global synchronization regime is
equal to the mean rotation number of all elements in the ensemble (as in the
case of coupled phase (Chap. 5) and limit-cycle (Chap. 6) oscillators).
For randomly distributed frequencies ωj , the evolution of the rotation num-
ber distribution is qualitatively quite similar to the case of continuous-time
active rotators (Chap. 5). The following three types of transitions can then be
observed in the synchronization trees (1) two adjacent elements (clusters) with
close frequencies can be easily synchronized and a new cluster appears; (2)
a regime of nonlocal synchronization can occur, i.e., an element (a cluster of
elements) becomes synchronized not to a nearest-neighbor element (cluster),
but to some other element (cluster) having a close rotation number. At that
the rotation numbers of the elements (clusters) in between are considerably
9.5 Chaotic Phase Synchronization 207
different; and (3) one element (group of elements) at the edge of one cluster
can go to another neighboring cluster.
Synchronization clusters are more stable (Fig. 9.11c, d) with respect to
the increase of noncoherence of rotations. Small c does not practically change
the number of clusters, the number of elements in the clusters, and mean
rotation numbers. With a further increase of the noncoherence, a transition to
the global synchronization regime through the appearance of well-pronounced
clusters is still observed. Only the structure of intermediate clustered states
can be different.
φkj+1 − φkj
sj = sin2 ( ) (9.44)
2
which characterizes the instantaneous phase difference between neighboring
oscillators. We have then that sj = 0 if the phases are equal and sj = 1 if
they differ by π. The space–time behavior of the boundaries between clusters
corresponds to the positions where phase difference slips or defects occur.
These defects are clearly seen as maxima (black regions) of sj . They can
follow regularly in time at certain positions in the chain; this case corresponds
to the existence of strong jumps between the clusters (Fig. 9.14c). If cluster
structures do not exist or the borders between them are smooth, then defects
appear irregularly in both space and time (Fig. 9.14d).
208 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
1.22 1.22
1.12 1.12
ρj /ρ1
1.10 1.10
1.08 1.08
1.06 1.06
1.04 1.04
1.02 1.02
1.00 1.00
0.98 0.98
0 10 20 30 40 50 0 10 20 30 40 50
j j
Fig. 9.13. Hard (a) and soft (b) transitions to regime of global chaotic phase syn-
chronization in system (9.1). Relative rotation numbers ρj /ρ1 for different coupling
coefficients d for linear distribution of individual frequencies and for ω1 = 0.6, fre-
quency mismatch Δ = 0.0002, c = −0.002 (a) and c = −0.4 (b)
The critical values of coupling corresponding to the onset of the global syn-
chronization regime in a chain of 50 chaotic CMs with a linear distribution
of individual frequencies ωj for ω1 = 0.6, different values of frequency mis-
match Δ and different values of c are presented in Fig. 9.15. With an increase
of the parameter −c the value d+ that slightly increases first and then can
decrease and increase again, and finally increases. After some critical value
−c∗ a synchronization is impossible due to the very high noncoherence of
rotations.
9.6 Conclusions
1600 3200
1200 2400
800 1600
400 800
k=0 k=0
1 j 50 1 j 50 1 j 50 1 j 50
Fig. 9.14. Space–time plots of evolution of (a, b) sin(φkj ) and sj (9.44) (c, d) by
hard (a, c) and soft (b, d) transitions to a regime of global chaotic phase synchroniza-
tion for a linear distribution of individual frequencies in (9.1). The parameters are
N = 50, ω1 = 0.6, frequency mismatch Δ = 0.002, coupling d = 0.39 and c = −0.002
(a, c) and c = −0.4 (b, d)
1. Identical maps
– The stability conditions depend only on the eigenvalue ρo of the indi-
vidual map, the number N of maps in the ensemble, and on the asym-
metry of coupling. For fixed ρ0 , there is a critical size Ncr of the chain
such that for N > Ncr a homogeneous synchronization regime cannot
occur for any coupling strength. For a fixed size N there is a critical
210 9 Regular and Chaotic Phase Synchronization of Coupled Circle Maps
0.9
Δ=0.0005
0.8 0.001
0.002
0.7
0.6
d+
0.5
0.4
0.3
0.2
0.1
− 0.8 − 0.7 − 0.6 − 0.5 − 0.4 − 0.3 − 0.2 − 0.1 0.0
c
Fig. 9.15. Critical value of coupling corresponding to the transition to global syn-
chronization regime vs. −c for different values of the frequency mismatch Δ in system
(9.1)
2. Nonidentical maps
– For chains of coupled circle maps, the typical features are the onset
and the existence of global (all-to-all) and cluster (partial) synchro-
nization regimes. These regimes have been observed both for regular
and chaotic maps.
– For high coherent rotations (c = 0) and for any type of frequency
distribution the rotation number of the global synchronization regime
is equal to the mean rotation number of elements in the ensemble.
– As well as for identical maps, increase of coupling strength can lead to
desynchronization phenomena, i.e., global or cluster synchronization
regime is changed by fully incoherent nonsynchronous state.
– As for chains of periodic and chaotic continuous in time oscillators,
for coupled maps two scenarios of transition to global synchronization
regime or transition between cluster structures have been found (1) a
gradual adjustment of the rotation numbers is observed and (2) the
transition occurs through the appearance of synchronized clusters.
– Hard transition between the cluster structures is more typical for
highly coherent rotations, while a soft transition is more often observed
for noncoherent rotations.
All presented properties especially the result that a synchronization regime
can be destroyed through increasing of the coupling strength is of special
importance for the design of DPLLs in order to realize stable synchronization
in engineering applications.
10
Controlling Phase Synchronization
in Oscillatory Networks
(a)
OSCILLATOR 1 OSCILLATOR 2
(b)
OSCILLATOR 1 OSCILLATOR 2
(c)
α1 α2
OSCILLATOR 1 OSCILLATOR 2
X1 X2
CONTROLLER
where x1,2 and F1,2 are n-dimensional vectors, ω1,2 are parameters defining the
time dependence rate (in some cases, frequencies) of oscillators x1,2 (t).1 Our
purpose is to synchronize two such oscillators by using a feedback control of
the timescales of coupled oscillators in such a way that the new characteristic
−1
time scales T1,2 ∼ Ω1,2 become identical. Here Ω1,2 are the mean observed
frequencies of the oscillators being controlled. In order to synchronize coupled
subsystems, we apply a feedback control in the following form (Fig. 10.1c):
dk dk−1 d
L = γk k
+ γk−1 k−1
+ ... + γ1 + a0 (10.3)
dt dt dt
1
Often, the time dependence rates (or frequencies) can be expressed in terms of
multipliers of the right-hand parts: ẋ1,2 = ω1,2 F1,2 (x1,2 ).
10.1 General Principles of Automatic Synchronization 215
1. The two signals x1 and x2 taken from both interacting systems go to a mul-
tiplier (first part of the “Controller” shown in Fig. 10.1c) which generates
the product u = Q(x1 , x2 ) of these signals. The spectrum of the oscilla-
tions Q consists of a “low” part defined by the difference Ω2 − Ω1 and a
“high” part defined by the sum Ω2 + Ω1 .
2. The signal u is conducted through the low-pass filter (second part of the
“Controller” shown in Fig. 10.1c), which damps the “high” frequency part
due to a specially designed transfer function. Hence, the control variable
u(t) becomes a slow-varying time function, whose spectral band goes to
zero. After the filtering, u(t) is added to both interacting systems (10.2) in
such a way that it may change their characteristic time scales. The main
goal is that this procedure provides a balance between the new timescales,
i.e., Ω2 = Ω1 .
In addition to the comparison of the observed frequencies of the controlled
systems, we are also interested in the evolution of their phase difference, which
is typically used in the study of PS.
This principle of getting synchronization is effectively used in applications
of PLL in a large number of radio- and telecommunication devices, radio-
location [57], coupled lasers [333,334], etc. It also takes place in a huge variety
of examples in nature, where the interaction of some oscillatory objects leads
to their balanced behavior. This balanced behavior is achieved by a nonlinear
interaction of the elements [8, 335–340]. Usually this coupling has the form of
a quadratic function of the interacting elements [346]. This type of coupling
is able to minimize efficiently the oscillator’s phase difference and, therefore,
it leads to synchronization.
It is important to emphasize that this principle can be applied not only
to coupled self-oscillatory systems but to other systems as well. For example,
let us consider a controlled linear oscillator:
ẋ = y,
ẏ = (ω 2 + αu)x + λy, (10.5)
τ u̇ = −u + βxy.
ρ̇ = λρ(p2 − ρ2 ),
φ̇ = (u + ω2 ), (10.7)
u̇ = −γu + βp cos(ω1 t + ψ1 )ρ cos φ.
where x1 , y1 , z1 are the variables of the Rössler oscillator and x2 , y2 are those
of the van der Pol oscillator. u is again the control variable added to both
subsystems, α1,2 , β and γ are the control parameters. We set β = γ = 1. For
the van der Pol oscillator we choose the following set of parameters: ω2 = 1,
ε = 0.01, and p = 4. The parameters of the Rössler oscillator will be chosen
as: a ∈ [0.15 : 0.2], b = 0.1, c = 8.5, and ω1 = 1.2 For these values the
topology of the chaotic Rössler attractor is rather simple, i.e., phase coherent
(see Sect. 2.3), and one can introduce the phase in the form
Because for the chosen ε the phase trajectory of the van der Pol oscillator
regularly monotonously oscillates around the origin, we can use a similar def-
inition of the phase
φ2 = − arctan(y2 /x2 ). (10.12)
In order to test for the existence of PS between Rössler and van der Pol
oscillators, we use as in the previous chapters the two criteria:
1. PS sets in if the mean frequencies of both coupled subsystems become
equal (frequency locking):
Ω2 = Ω 1 , (10.13)
where the frequencies are defined as
φ1,2 (T ) − φ1,2 (0)
Ω1,2 = lim . (10.14)
T →∞ T
2
For ω1 = 1 in the chaotic Rössler oscillator the mean in time frequency does not
equal to 1 and therefore, uncoupled oscillators have some frequency mismatch.
218 10 Controlling Phase Synchronization in Oscillatory Networks
λ, Ω1−ω2 0.1
−0.1
−0.3
−0.5
0.000 0.001 0.002 0.003 0.004
18.0
14.0
x1max
10.0
6.0
2.0
0.000 0.001 0.002 0.003 0.004
α1
We have also performed numerical simulations where the van der Pol and
the Rössler oscillator are mutually coupled by feedback (α1,2 = 0). The effect
of both regular and chaotic PS has been observed there as well.
220 10 Controlling Phase Synchronization in Oscillatory Networks
where x1,2 , y1,2 , z1,2 are the variables of the first and second Rössler oscillator,
respectively. We set a = 0.15, b = 0.1, c = 8.5, ω1 = 0.98, and ω2 = 1.02.
Hence, for both oscillators the phase definitions (10.11) can be used. The
10.4 Two Coupled Rössler Oscillators 221
0.2
λ,Ω1−Ω2
0.1
0.0
−0.1
0.0000 0.0002 α1 0.0004 0.0006
250
200
α1=0.0004
150
φ2−φ1
100 α1=0.000405
50 α1=0.00041
α1=0.000415
0
−50
0 20000 40000 60000 80000 100000
time
Fig. 10.4. Synchronization of two coupled Rössler oscillators (10.17). The parame-
ters are a = 0.15, b = 0.1, c = 8.5, ω1 = 0.98, ω2 = 1.02, α1 = −α2 , and β = γ = 1.
(a) The four largest LEs and the difference of the mean observed frequencies Ω1 −Ω2
(circles) vs. the control parameter α1 . (b) Difference of the phases φ2 − φ1 for
nonsynchronous (α1 = 0.0004; 0.000405; 0.00041) and synchronous (α1 = 0.000415)
regimes
222 10 Controlling Phase Synchronization in Oscillatory Networks
The equation for the phase difference of (10.18) is (for details see Sect. 4.2.1)
β ρ21 + ρ22
θ̇ − sin θ = Δ (10.22)
2 ρ1 ρ2
The stable stationary state
1.05
1.04
a = 0.15
1.03
Ω2/Ω1
a = 0.16
1.02 a = 0.17
a = 0.18
1.01
1.00
0.99
0.00 0.02 0.04 0.06
α1
Fig. 10.5. Synchronization of two coupled Rössler oscillators (10.17). The ratio of
the mean observed frequencies Ω2 /Ω1 vs. the control parameter α1 . The parameters
of the individual oscillators are the same as in Fig.10.4
where x1,2 , y1,2 , z1,2 are the variables of the Rössler and Lorenz oscillators,
respectively. The parameters a, b, c and phase of the Rössler oscillator are the
same as in the previous case; ω = 0.98, τ = 8.3, σ = 10, r = 28, and b = 8/3
and the phase of the Rössler oscillator is measured as before. The phase of
the Lorenz oscillator is calculated as θ = arctan((z2 − 27)/( x22 + y22 − 12))
(see Sect. 2.4). In Fig. 10.6 we present results of the transition to chaotic PS
between Rössler and Lorenz oscillators. One can see an interval of α where
PS occurs. Therefore, using the proposed scheme we are also able to achieve
chaotic PS between oscillators with a strong difference in their topology.
224 10 Controlling Phase Synchronization in Oscillatory Networks
0.00
− 0.05
Ω1−Ω2
− 0.10
− 0.15 γ = 0.6
0.8
− 0.20 1.0
1.2
− 0.25
0.0 0.0005 0.001 0.0015 0.002
α
Fig. 10.6. Synchronization of coupled Rössler and Lorenz oscillators (10.27). The
parameters are a = 0.15, b = 0.1, c = 8.5, ω = 0.98, σ = 10, r = 28, b = 8/3, α1 =
−α2 = α, and β = γ = 1. The difference of the mean observed frequencies Ω1 − Ω2
vs. the control parameter α
where xj and Fj are n-vectors, ωj are parameters defining the time dependence
rate of the oscillations xj (t), and N is a number of oscillators.
In order to synchronize these systems, we generalize the feedback control
(10.2) in the following form:
ẋj = Fj (xj , ωj + αj uj ),
Luj = Qj (x1 , . . . , xN ), (10.29)
j = 1, . . . , N,
N
Qj (x1 , . . . , xN ) = Qk (xj , xk ), (10.30)
k=1,k=j
CONTROLLER CONTROLLER
j−1 j
Now we study whether the control variable uj (t) added to each oscillator
can provide a synchronous behavior between interacting elements. Figure 10.7
presents a simple scheme which roughly describes the proposed coupling tech-
nique (cf. also the scheme of a multichannel chain of partial digital phase-
locked loops connected in parallel by phase-mismatching signals, Fig. 9.1).
We demonstrate the method of feedback control for PS for ensembles of
(1) locally coupled regular oscillators (Sect. 10.7), (2) locally coupled chaotic
oscillators (Sect. 10.8), and (3) globally coupled chaotic oscillators (Sect. 10.9).
Δ
Ωs = ω1 + (N − 1) (10.40)
2
Then the frequencies for all elements are equal to the mean frequency of
the elements in the ensemble. With an increase of the frequency mismatch Δ
(or decrease of the coupling α), a loss of global synchronization regime takes
place. For a long chain the two synchronization clusters occur, i.e., the chain
is divided into two clusters each of size (N/2) both consisting of mutually syn-
chronized oscillators. Further increase of Δ (decrease of α) leads to a sequence
of destruction of the one cluster structure of the synchronized elements and
to the appearance of another structure. This sequence obtained in numerical
experiments is shown in Fig. 10.8. From this figure we recognize two types of
transitions between cluster structures. In the first type a hard transition with-
out intermediate structures occurs from the state with n (n + 1) clusters to
the state with n + 1 (n) clusters (see, for example, the interval [0.019 : 0.023]).
In the second type, a soft transition happens with a smooth transition of
intermediate structures one into the other. As follows from Fig. 10.8, the the-
oretically and numerically obtained conditions of the global synchronization
regime and the frequency of the global synchronization regime are in very
good agreement.
1.020
1.018
1.016
1.014
1.012
1.010
Ωj
1.008
1.006
1.004
1.002
1.000
0.998
0.00 0.01 0.02 0.03 0.04 0.05 0.06
α
Fig. 10.8. Observed frequencies Ωj in a chain of Poincaré systems (10.31) with
linear distribution of individual frequencies vs. α. N = 20, p = 1, ω1 = 0.98,
Δ = 0.001
228 10 Controlling Phase Synchronization in Oscillatory Networks
ẋj = −ωj yj − zj
−ωj α(sin(φj+1 − φj ) + sin(φj−1 − φj ))yj ,
ẏj = ωj xj + ayj (10.43)
+ωj α(sin(φj+1 − φj ) + sin(φj−1 − φj ))xj ,
żj = b − czj + xj zj ,
j = 1, . . . , N.
1.04
1.035
Ωj
1.03
α = 0.0
0.04
0.1
0.16
0 20 40 60 80 100
j
1.06
α=0.0
0.01
1.055 0.02
0.06
1.05
1.045
Ωj
1.04
1.035
1.03
1.025
0 20 40 60 80 100
j
We take the same parameters of individual elements as in Sect. 10.8 and ran-
domly distributed frequencies ωj . Let us fix again all γj 1. Then the filtered
control variable uj can be described in the form:
N
uj = sin(φj − φk ) (10.45)
k=1,k=j
1.0
0.8
0.6
<R>
0.4
0.2
0.0
0.0 0.000005 0.00001 0.000015 0.00002
α
Fig. 10.11. Frequency entrainment in the ensemble of globally feedback coupled
Rössler oscillators (10.46) with randomly distributed frequencies in the interval
[0.98,1]. Order parameter R (10.47) vs. coupling α. The number of elements
N = 100
10.10 Conclusions 231
ẋj = −ωj yj − zj
N
−ωj α k=1,k=j sin(φj − φk )yj ,
ẏj = ωj xj + ayj
N (10.46)
+ωj α k=1,k=j sin(φj − φk )xj ,
żj = b − czj + xj zj ,
j = 1, ..., N.
10.10 Conclusions
In this chapter we have described a feedback control method for automatic
phase locking of regular and chaotic oscillators.
The main advantages of this method are the following:
This presented approach can be helpful (1) for the understanding of self-
organization mechanisms in many systems in nature and possibly also for
their manipulation and (2) for the design of different schemes of automatic
synchronization and could be applied to communication, engineering, ecology,
and medicine.
232 10 Controlling Phase Synchronization in Oscillatory Networks
Chain B :
A
a1 a2 a3 aN
B
b1 b2 b3 bN
0.8
V 0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
d1
Fig. 11.2. Velocity v of the front of the transition from the excited to the non-excited
state in (11.1) as a function of the coefficient of coupling between the cells inside
the chain (Δ = 2.0, α = 5.75, c = 0.51). “+” correspond √ to numerical simulations.
Dashed line corresponds to the dependence v ≈ 0.92 d1 − 0.09
At weak coupling between the elements inside the array (d1 dcr 1 = 0.09),
a motionless (stationary) front that separates the oscillators into excited states
and non-excited ones is formed. At strong coupling (d1 > dcr 1 ), the regime of
oscillator death prevails, and the region in which this is realized broadens and
embraces all the array. The transition process and the influence of it on the
initial conditions occur simultaneously in a complicated way with a special
complication of the structures formed near the critical value dcr 1 ≈ 0.09. But
the velocity v of the front propagation of the oscillator death may still be
determined for some realizations, i.e., at a definite choice of initial conditions.
Numerical simulation shows that the velocity remains almost unchanged in
the range of the chain coupling c = [0.5, 1.0] and has a root dependence on
cr
√ parameter at d1 > d1 as it is typical for critical phenomena:
the coupling
v ≈ 0.92 d1 − 0.09 (Fig. 11.2). It is worth to note that a deviation from the
root dependence and the jump-wise transition from zero to nonzero velocities
at nonzero coupling coefficients d inside the chains is observed only in discrete
models. Partial differential equations for complex amplitudes predict a smooth
transition starting from dcr1 = 0.
A more detailed analysis of the transition processes at the interface
between both regions having two different states gives a direct confirmation
that the transition to an unexcited state is actually determined by desyn-
chronization. This is illustrated in Fig. 11.3 by time series of the real and
imaginary parts of the complex amplitudes aj and bj on one arbitrary chosen
cell (j = 100) in each chain at the moving front. Another useful characteristic
is the difference of phases ϕa and ϕb in both chains. The damping of the oscil-
lations is preceded by an abrupt change in the phase difference ϕa100 − ϕb100
with a subsequent formation of oscillations having different frequencies which
11.3 Dissipative Coupling (Zero “Dispersion”) 237
1.0
|a100|2 (a)
0.8 s
0.6
0.4
0.2
0.0
100 150 200 250 300
1.0
Re(a100)
(b)
0.5 Re(b100)
0.0
− 0.5
−1.0
100 150 200 250 300
t
Fig. 11.3. Desynchronization in (11.1) and (11.2). Plot of the oscillation intensity
ϕa −ϕb
(|a100 |2 ) and of s = sin2 ( 100 2 100 ) characterizing the phase difference (a) and real
parts of complex amplitudes a100 and b100 (b) of the 100th oscillator at the transition
to the trivial equilibrium state aj = bj = 0 (Δ = 2.0, α = 5.75, p = 0.5, c = 0.51,
d1 = 0.15, N = 128)
Another typical feature in the system (11.1) and (11.2) is the existence of
localized synchronization structures. The bistability of the stationary regimes
allows us to chose initial conditions in such a way that these structures can
be formed in any place of the chain. Fig. 11.4 presents the structures that are
formed when localized oscillations aj = bj = |a| with amplitudes determined
according to (11.3) are specified at the initial instant of time. The distributions
of initial and established amplitudes almost coincide at weak coupling (d1
dcr
1 ) between the oscillators inside the chains. As the internal coupling d is
increased, the influence of the coupling with unexcited oscillators becomes
significant. As a result the amplitude |aj | and, consequently, the nonlinear
phase shift determined by the term α|aj |2 aj strongly depend on their position
in the chain. At d1 ≈ dcr 1 , this shift increases so that a local breakdown of
238 11 Chains of Limit-Cycle Oscillators
0.5
d1=0.001
0.05
0.4 0.1
0.3
|aj|2
0.2
0.1
0.0
0 16 32 48 64 80 96 112 128
j
Fig. 11.4. Stationary structures of system (11.1) and (11.2) with parameters Δ =
2.0, α = 5.75, p = 0.5, c = 0.51, and d = 0.001, 0.05, 0.1 that are realized under the
initial conditions a2j = b2j = 0 for j = 1, . . . , 50, a2j = b2j = 0.434 for j = 51, . . . , 78,
and a2j = b2j = 0 for j = 79, . . . , 128
the synchronization regime may occur and smaller structures are formed. The
resulting amplitude distribution near this critical value depends significantly
on the coupling and the initial amplitude.
0.5
φ14−φ19
−0.5
1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000
t
Fig. 11.5. Nonlocal synchronization regime in (11.1) and (11.2). Onset of in-phase
synchronization regime at different initial conditions in two isolated clusters (I for
j = 13, 14 and III for j = 19, 20) separated by cluster II for j = 16, 17 for the chains
with parameters N = 32, p = 1.0, α = 8.0, c = 2.0, d = 0.15, and Δj = 6.0 for
j = 1, . . . , 15, and j = 18, . . . , 32 and Δj = 5.9 for j = 16, 17
for c in (11.1) and (11.2), respectively). Supposing that the oscillatory struc-
ture is homogeneous (aj = a, bj = b), stable regimes of pulsations and
nonsynchronized oscillations are possible besides the regimes of synchroniza-
tion and oscillator death. Clusters of oscillations in one of these regimes,
depending on the initial conditions, are formed at sufficiently weak coupling
d1 . When one of the competing regimes is synchronized and the other one
is not, significantly nonstationary structures may be formed in the system,
which leads to a specific turbulent regime called spatiotemporal intermittency
(Fig. 11.6a) (see also transition to spatiotemporal intermittency in chains of
(a) (b)
1000
800
600
400
200
t=0
1 50 100 1 50 100
j j
coupled intermittent chaotic maps described in Chap. 8). In this case, the
front propagation becomes chaotic due to a non-stationary dynamics of the
domains, giving rise to a process similar to directional percolation [369].
As the coupling d1 between the cells is increased, only those structures
which consist of a small amount of oscillators with strong oscillations survive,
like in the case of symmetric coupling between the chains. At a small degree of
asymmetry of coupling between the chains, intensive oscillations are excited
in isolated groups of oscillators and occur as short trains. The position of such
oscillations is random in space and time. As the nonsymmetry of coupling is
increased, the average density and repetition rate of these trains grow and
there occurs a transition from oscillations intermittent in space and time to
developed turbulence – spatiotemporal intermittency (STI) (Fig. 11.6b).
25000
t=0
j=1 100
Fig. 11.7. Bursting structure in (11.1) and (11.2). Dependence of the oscillation
intensity |aj (t)|2 on the element’s position in the array j and time t (shading corre-
sponds to the intensity)
70
60
number of bursts
50
40
30
20
10
0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95
d2
0.4 (a)
|a(90)|2
0.3
0.2
0.1
0
0 5000 10000 15000 20000
time
0.4
|a(90)|2
0.3 (b)
0.2
0.1
0
0 5000 10000 15000 20000
time
Fig. 11.9. Examples of time series showing intermittency in (11.1) and (11.2).
Intervals of laminar behavior are randomly interrupted by short bursts: (a) d2 = 0.56
and (b) d2 = 0.72
0.190
0.185
d2=0.55
Ai+1 d2=0.547
d2=0.546
0.180
0.175
0.175 0.180 0.185 0.190
Ai
Fig. 11.10. The first return map constructed from the maxima of the time series
of the intensity of oscillations of the 95th element in (11.1) for three different values
of d2 : d2 = 0.546 – before the bifurcation (stable attracting point), d2 = 0.547
and d2 = 0.55 – after the bifurcation from the quasistationary mode (quadratic-like
map); d1 = 0.3
250
200
150
N
100
50
0
200.0 1200.0 2200.0
L
Fig. 11.11. Histogram for the distribution of laminar intervals length for the 95th
element in (11.1) after the bifurcation from the quasistationary mode. Parameters:
d1 = 0.3, d2 = 0.55
0.210
0.190
Ai+1
d2=0.75
d2=0.72
0.170
0.150
0.150 0.170 0.190 0.210
Ai
Fig. 11.12. The first return map constructed from the maxima of the time series
of the intensity of oscillations of the 95th element in (11.1) for two different values
of d2 after bifurcation from the stationary mode, d1 = 0.302
125
100
75
N
50
25
0
0.0 2000.0 4000.0 6000.0
L
Fig. 11.13. Example of a histogram for the distribution of laminar interval length
for the 95th element in (11.1) after the bifurcation from the stationary mode. Para-
meters: d1 = 0.302, d2 = 0.75
11.4 Nonscalar (Dissipative and Conservative) Coupling 247
strictly one dimensional. The examination of the bifurcation suggests that the
bifurcation may correspond to type-I intermittency (i.e., the Floquet multi-
plier crosses the unit circle at +1), but the return map has a significant high-
order (for example, cubic) term, i.e., this term is essential near the bifurcation
point (a similar map with an essential cubic term was considered in [376]).
Chain B :
140
120
number of bursts
100
80
60
40
20
0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95
d2
Fig. 11.14. The number of bursts per 5 × 104 times units in the system with noise
in (11.6) and (11.7) vs. d2 , d1 = 0.302. The amplitude of noise δ = 0 (no noise, +),
0.0001 (×), 0.0005 (∗), 0.001 (), 0.002 (), and 0.003 (◦). The curves break at the
values of δ, where intermittency is replaced by pure propagation
248 11 Chains of Limit-Cycle Oscillators
the noise intensity δ = 0.002 and 0.003, a chaotic bursting arises for such
values of d2 , where bursting is observed. The most important observation is,
however, that the number of chaotic bursts decreases and even vanishes for
certain domains of d2 . Thus, noise can suppress chaotic intermittent bursting
and provides a way to control the intermittent front dynamics. For a possible
explanation of this nontrivial effect, the multistability of different regimes
should be taken into account. Under the noise influence, the bursting regimes
become unstable and only the trivial state (aj = bj = 0) remains attracting.
11.5 Conclusions
In this chapter, effects of collective oscillations in a system of two coupled
chains of identical limit-cycle oscillators, namely coupled discrete Ginzburg–
Landau equations, are investigated. The main findings are the following:
– A pair of coupled chains with different collective frequencies exhibit stable
fronts between two possible asymptotic states: synchronous oscillations
and oscillation death
– The inhomogeneous states formed by the fronts persist at weak coupling
between the oscillators in each chain due to discrete space variable; thus
providing conditions for the existence of localized synchronization struc-
tures
– At stronger coupling, the interface between the two regions of the chains
may propagate
– Synchronized clusters can be induced by disorder
– Transitions from nonpropagating to propagating fronts between synchro-
nized and nonexcited oscillators occur via intermittency
– Noise plays a constructive role in regularization of collective behavior of
oscillators
On the Basis of these finding we would like to underline the following: The
mechanisms of the formation of localized structures considered above have
a strong impact on a better understanding of the origin of low-dimensional
chaos in multidimensional and extended systems. In particular, weakly cou-
pled clusters with m < N cells are frequently formed in large ensembles of
N oscillators, including the ones having different parameters. As to expect
from the above analysis, only clusters with synchronized (at least partially)
cells may survive in a definite region of the parameters. As a result, the
effective number of degrees of freedom sufficient for the description of the
dynamics of the system decreases substantially. Our analysis confirms that,
at certain conditions, this mechanism is also effective at chaotic synchroniza-
tion (see Chaps. 7–9) and leads to the formation of localized structures with
low-dimensional chaotic dynamics. On the other hand, the formation of vari-
ous complex patterns observed at relatively simple but inhomogeneous initial
conditions indicates that the studied phenomenon may also be responsible
11.5 Conclusions 249
12.1 Objectives
The forms of all these currents are given in Sect. 2.7.3. gi (V, t) is the product
of one or more gating variables, g∞ = αgi /[αgi + βgi ] is the steady-state value,
τgi = 1/[αgi + βgi ] is the time constant, and the αs and βs are the functions
of the membrane voltage. The conductances and reversal potentials used in
the simulations are listed in Sect. 2.7.3.
15
1:1
10
Output Freq. (Hz)
2:1
0
0 5 10 15 20
Input Freq. (Hz)
Fig. 12.1. Output frequency vs. input frequency for a driven one-dimensional chain
of 30 cells. The stimulus consists of a square wave with an amplitude of 100 μA cm−2
and a duty cycle of 50%. The input–output frequency characterization for non-
excited cells, initial conditions are marked by plus and for adapted initial conditions
by circle. The dashed horizontal line at 6.3 Hz represents the spiral wave frequency
of cardiac tissue
110
90
Amplitude 70
50
30
10
−10
0 T 2T 3T 4T 5T
time
Fig. 12.2. External force as a sequence of rectangle pulses with a 50% (T /2) duty
cycle
10 10
5 5
0 0
0 10 20 30 0 10 20 30
10 10
5 5
0 0
0 10 20 30 0 10 20 30
10 10
5 5
0 0
0 10 20 30 0 10 20 30
Input Freq. (Hz) Input Freq. (Hz)
Fig. 12.3. Output frequency vs. input frequency for a driven one-dimensional chain
of 30 cells. The stimulus consists of a square wave (see Fig. 12.2) with an amplitude
of 100 μA cm−2 and a duty cycle of: (a) a fixed, 2 ms duration; (b) 12.5%; (c) 25.0%;
(d) 37.5%; (e) 50.0%; and (f ) 62.5%. As in Fig. 12.1, the dashed horizontal lines
represent the intrinsic spiral wave frequency of 6.3 Hz. Note that the part of Fig. 12.1
has been included as (e) for reference. Initial conditions: non-excited cells
region is invaded by surrounding 2:1 synchrony regions. This clarifies why the
1:1 intermittency region quickly disappears if the duty cycle departs from a
value of 50%.
We also examine effects of stimulus waveform on the 1:1 and 2:1 synchrony
regions. Figure 12.6 presents the results for six different waveforms typically
used for cardiac pacing [395]:
1. Monophasic square-wave pulses
2. Symmetric biphasic square-wave pulses of 2 ms duration
3. Sinusoidal force
4. Pulses with ramp wavefronts and vertical waveback
5. Pulses with vertical wavefronts and ramp waveback
6. Pulses with ramp wavefronts and waveback
12.4 Computational Results 259
1:1 synchrony
2:1 synchrony
100
Stimulus Ampl. (μAmp/cm2)
80
60
40
20
6 8 10 12 14 16 18 20
Input Freq. (Hz)
Fig. 12.4. Regions of 1:1 plus and 2:1 circle synchrony, as a function of stimulus
amplitude and input frequency, for a driven one-dimensional chain of 30 cells. The
stimulus consists of a square wave with a duty cycle of 50%. Initial conditions: non-
excited cells. Middle region marked by plus is 1:1 intermittency synchrony region
1:1 synchrony
2:1 synchrony
Stimulus Ampl. (μAmp/cm2)
20 20
10 15 20 10 15 20
20
10 15 20
Stimulus Ampl. (μAmp/cm2)
20 20
10 15 20 10 15 20
Input Freq. (Hz) Input Freq. (Hz)
Fig. 12.5. 1:1 plus and 2:1 circle synchrony regions, as a function of stimulus
amplitude and input frequency, for a driven one-dimensional chain of 30 cells. The
stimulus consists of a square wave with a duty cycle of: (a) 40%, (b) 45%, (c) 50%,
(d) 55%, and (e) 60%. Note that Fig. 12.4 has been included as c for reference.
Initial conditions: non-excited cells
keeping the membrane voltage elevated. This would make a cell less responsive
to high-frequency inputs.
We also investigate the effects of the calcium channel antagonists on our
one-dimensional cardiac model. Therefore, we eliminate transmembrane cal-
cium fluxes by setting Gsi = 0 throughout each in numero experiment. The
input–output frequency response for various duty cycles are shown in Fig. 12.7.
The main results are:
1. The elimination of slow inward calcium currents serves to increase the
intrinsic spiral wave frequency from 6.3 to 25.0 Hz.
2. The calcium channel antagonists eliminate the intermittency regions found
in the original model (Fig. 12.7). Hence, we can conclude that it may
only be possible to eliminate spiral waves using a combined action of
calcium channel blockers and overdrive pacing with a narrow band of
input frequencies around 27 Hz at duty cycles less than or equal to 37.5%.
Our simulation shows the absence of bistability.
12.4 Computational Results 261
1:1 synchrony
2:1 synchrony
20 20
10 15 20 10 15 20
Stimulus Ampl. (μAmp / cm2)
20 20
10 15 20 10 15 20
20 20
10 15 20 10 15 20
Input Freq. (Hz) Input Freq. (Hz)
Fig. 12.6. 1:1 plus and 2:1 circle synchrony regions, as a function of stimulus
amplitude and input frequency, for a driven one-dimensional chain of 30 cells. The
stimulus consists of the following waveforms (shown to the right of each plot): (a) a
monophasic square-wave pulse of 2 ms duration, (b) symmetric biphasic square-wave
pulses of 2 ms duration, (c) a sine wave, (d) a waveform with a ramp wavefront and
a vertical waveback, (e) a waveform with a vertical wavefront and a ramp waveback,
and (f ) a triangle wave. Initial conditions: non-excited cells
We use the results from the one-dimensional simulations to explore the possi-
ble suppression of spiral waves and spiral wave chaos in the practically much
more interesting case of two dimensions. The analysis of the one-dimensional
chain in Fig. 12.4 suggests especially that overdrive pacing of the system at a
frequency of 12.4 Hz with a square-wave stimulus of amplitude 50 μA cm−2
and duty cycle 50.0% should be sufficient for suppression of spiral waves
in a two-dimensional medium. First, we present sequential snapshots of the
two-dimensional cardiac model subject to such stimulation (Fig. 12.8). The
excitation was applied, starting at t = 0.00 s, with a model equivalent to a
strip electrode at the top of the medium to ensure a fast conduction velocity.
Figure 12.8 also shows that the initial spiral wave chaos (t = 0.00 s) is entirely
eliminated after 5.00 s of pacing. By continuing the pacing beyond this point,
we see that more-or-less planar wavefronts travel through the medium for at
262 12 Chains and Lattices of Excitable Luo–Rudy Systems
20 20
10 10
20 30 40 20 30 40
30 30
20 20
10 10
20 30 40 20 30 40
20 20
10 10
20 30 40 20 30 40
Input Freq. (Hz) Input Freq. (Hz)
Fig. 12.7. Output frequency vs. input frequency for a driven one-dimensional chain
of 30 cells. The calcium current is eliminated in the model by setting Gsi = 0
throughout each in numero experiment. The stimulus consists of a square wave with
an amplitude of 100 μA cm−2 and a duty cycle of: (a) a fixed, 2 ms duration; (b)
10.0%; (c) 12.5%; (d) 25.0%; (e) 37.5%; and (f ) 50.0%. The dashed horizontal lines
represent the intrinsic spiral wave frequency of 25.0 Hz
least 1 s (from t = 5.00 s to t = 6.00 s). Shortly after, however, spiral wave
chaos is reinitiated, as can be seen at t = 6.25 s. Thus, there exists a window,
in this case of approximately 1.0 s, in which the medium is free of any spiral
wave chaos. If the pacing is halted during this window, the medium will be
left in a resting state. This finding has important consequences for cardiology
and shows especially that controlled intervals of overdrive pacing can be used
to suppress spiral wave chaos.
In general, however, we get limited success for suppressing spiral wave
chaos in two-dimensional media using overdrive pacing alone. The high-
frequency stimuli predicted from the one-dimensional studies did not always
result in planar wavebacks during capture. In many cases, subsequent wave-
fronts approach the wavebacks of prior stimuli, encountering cells in various
refractory states. This leads, in some instances, to the reinitiation of spiral
wave chaos after local capture. The main reason for this effect is the bistability
of the synchronous response of the media to the external stimulation.
12.4 Computational Results 263
Fig. 12.8. Sequential time images of a successful suppression of spiral wave chaos,
using overdrive pacing, in a network of 300×300 cells. The system is excited with the
model equivalent of a strip electrode of 2×300 cells located at the top border of the
medium. The stimulus consists of a square wave with an amplitude of 50 μA cm−2
and a duty cycle of 50%. The input frequency is 12.4 Hz. The membrane voltage
(mV) of the cells in each image is color coded as indicated in the bar located to the
right of the top row of panels
We inhibit the calcium flux in the model by setting the maximum cal-
cium conductance, Gsi , to zero after the first time step in the simulations.
Figure 12.9 presents results of a computational experiment where the top two
rows of cells were paced at a frequency of 27.9 Hz with a square-wave stim-
ulus of amplitude 50 μA cm−2 and duty cycle 12.5%. Inhibiting the calcium
flux causes the initial spiral wave chaos to conform to a more regular, quasi-
periodic meander, which is consistent with previously reported results [431].
This dynamic effect allows the applied stimulation to entrain the medium
progressively. After 4.0 s, the applied stimulation has already annihilated all
spiral wave activity within the medium (Fig. 12.9). When the stimulation is
subsequently halted, the medium returns to its resting state within 270 ms.
As was mentioned for the one-dimensional case, the synchronous response of
Fig. 12.9. Sequential time images of a successful suppression of spiral wave chaos,
using overdrive pacing in conjunction with calcium channel antagonists, in a network
of 300×300 cells. The system is excited with the model equivalent of a strip electrode
of 2×300 cells located at the top border of the medium. The stimulus consists of a
square wave with an amplitude of 50 μA cm−2 and a duty cycle of 12.5%. The input
frequency is 27.9 Hz. The calcium channels in the model are blocked at t = 0.0 s
by setting Gsi = 0. The membrane voltage (mV) of the cells in each image is color
coded as indicated in the bar located to the right of the top row of panels
12.5 Conclusions 265
the medium to the external stimulation is unique, i.e., it does not depend on
the current state of the cells in the media.
Next, we investigate effects of the electrode geometry on the suppression
of spiral waves by using a point source to generate convex waves. As it may be
impossible to generate a completely planar wavefront outside of the computa-
tional realm, it is important to investigate various electrode geometries. The
experimental conditions are similar to those in Fig. 12.9. We find that spiral
wave chaos could be suppressed; however, the time to get suppression is now
longer. We further explore the sensitivity of the spiral wave suppression to the
magnitude of the calcium current by repeating the above computational exper-
iment with modified values of the maximum calcium conductance, Gsi . Then,
spiral wave chaos could be suppressed similarly to that shown in Fig. 12.9
with reduced, nonzero levels of inward calcium current (Gsi greater than 0
but less than the nominal 0.07). However, the time to get suppression of spi-
ral wave chaos increases as the level of inward calcium current is increased.
In general, we yield more episodes of a successful suppression of spiral wave
chaos using the combined action of overdrive pacing and calcium channel
antagonists, than in using overdrive pacing alone. In fact, we find that spi-
ral wave chaos could be suppressed in all cases when the overdrive pacing is
coupled with calcium channel blockers.
12.5 Conclusions
θ̇ = Δω − ε sin θ, (13.1)
d
θ̇ = − V (θ), (13.2)
dθ
13.1 Degrading Effects of Noise: Noise-Induced Phase Slips 271
The ξ1 and ξ2 in the two oscillators are independent Gaussian noises with
intensity D. Introducing the phase φj by
x˙j
φj = − arctan , (13.6)
xj
0 60
(a) (b)
50
−1 40
V(θ)
30
θ
−2 20
10
−3 0
0 10 20 30 0 5000 10000
θ time
Fig. 13.1. (a) Systematic plot of the washboard potential V (θ) = −θΔω − ε cos θ
for the system (13.1). (b) Noise makes the phase difference θ fluctuate and induces
phase slips
272 13 Noise-Induced Synchronization
ΔΩ D=0.015 D=0.1
0.00 0.00
0.01 0.01 (d)
(b)
ΔΩ
D=0.05 D=0.3
0.00 0.00
0.017 0.018 0.019 0.020 0.021 0.017 0.018 0.019 0.020 0.021
ε ε
Fig. 13.2. Frequency difference ΔΩ vs. coupling strength ε. Solid lines: noise-free
case D = 0; circles: independent noise; Left panel (a,b): two coupled periodic van der
Pol oscillators (13.5); right panel (c,d): two coupled chaotic Rössler oscillators (13.7).
The natural frequencies are ω1 = 0.99 and ω2 = 0.97
we can compute the frequencies Ωj = φ̇j of the oscillations. Figure 13.2 (left
panel) shows the difference of the mean frequency ΔΩ = |Ω1 − Ω2 | as a
function of the coupling strength ε for various noise levels. Without noise,
D = 0, synchronization is achieved when ε > εps = Δω = 0.02, as predicted
by (13.3) (see the solid lines in Fig. 13.2a,b). When a small noise is included,
D = 0.015, we can see that the transition to synchronous regime is smeared
(Fig. 13.2). The frequency difference ΔΩ is not vanishing when the coupling
strength is slightly beyond the threshold εps , because noise induces phase slips.
These phase slips occur less frequently when the coupling becomes stronger,
so that the synchronization region is restored at larger coupling strength.
When the noise intensity is higher, phase slips happen more frequently, so that
effective synchronization with small ΔΩ ≈ 0 would require a much stronger
coupling strength, as seen in Fig. 13.2. Hence noise leads only to degrading
effects in periodic oscillators.
A different response to noise occurs in two coupled chaotic Rössler
oscillators
with ω1 = 0.99 and ω2 = 0.97. As seen in Fig. 13.2c,d by the solid line, the
transition point to the perfect PS regime in the absence of noise, εps = 0.0208,
is somewhat higher than that of the limit-cycle oscillators at εps = 0.020. In
the presence of noise, phase slips occur in the originally phase synchronized
regime, so that ΔΩ is not vanishing for ε slightly beyond the threshold. Com-
pared to the periodic oscillators in the left panel, it is important to emphasize
that noise makes the frequency difference ΔΩ smaller than that of the noise-
free oscillators for ε < εps . A smaller frequency difference indicates that phase
13.2 Noise-Induced CS and PS in Uncoupled Chaotic Oscillators 273
slips are reduced and a degree of synchrony is enhanced due to the noise, i.e.,
noise induces more order here.
This comparison shows that, in chaotic systems, noise may play a con-
structive role to induce and enhance synchronization. In the following, we
will show that noise can enhance synchrony of chaotic systems in different
ways and with different mechanisms. Similar effects we have also discussed in
Chap. 11.
Synchrony of nonlinear systems which are not coupled directly, but influ-
enced by a common random force, has drawn considerable attention and has
been a topic of a long-standing controversy. Firstly, it was shown that noise
can induce synchrony in periodic oscillations [441–443]. More interesting, the
phenomenon of noise-induced order was reported on a chaotic map which
is directly connected to the Belousov–Zhabotinsky chemical reaction [444].
There, a small amount of noise may change a chaotic trajectory of the sys-
tem into a state similar to a periodic orbit smeared with noise [444], which
makes the largest Lyapunov exponent (LE) (λ1 ) negative, and leads to a slower
decay of correlations and an improvement of the state predictability [445,446].
A negative largest LE means that, in an ensemble of systems with identical
laws of motion and common noise, such as the motion of floating particles on
a surface of an incompressible fluid [447], the states in the phase space shrink
into a single point [447, 448], i.e., noise induces CS in chaotic systems.
However, whether in general common noise can induce synchrony of
chaotic systems has been a topic of long-standing dispute. In [449], numer-
ical simulations show that a white noise can induce synchrony of the logis-
tic map, and in the Lorenz system, CS was observed for uniform noise in
[0, W ] with large enough W values, but not for unbiased one. In [450] it was
pointed out that the largest LE of the noisy logistic map is positive. The
CS observed in [449] is actually an artifact of finite precision in numerical
simulations [450, 451]. Several other authors, on the other hand, restudied
the problem by examining the properties of the noise applied to the systems.
The largest LE of the noisy Lorenz system is found to be the same as that
of the system driven constantly only by the mean value W/2 of the noise,
indicating that the bias of the noise plays the crucial role in synchroniza-
tion processes [452, 453]. Recently, in [454] authors analyzed CS of chaotic
systems by noise in an experiment with Chua’s circuit, drawing the general
conclusion that CS might be achieved only by a biased noise, but not by an
unbiased one, and synchrony is only a result of noise-induced order because
the biased noise drives the system into a noise-smeared periodic orbit, similar
274 13 Noise-Induced Synchronization
1
N
λ = lim ln |f (xn )| (13.11)
N →∞ N
n=1
becomes negative. Note that the Jacobian f (x) also governs the linearized
dynamics in the noise-free case. In this sense, the LE of stochastic systems is
well defined as in the deterministic systems [458]. The difference is that the
attractor in the phase space is modified by the noise. The noise-free map is
chaotic, i.e., λ > 0. To make λ < 0, we need a significant contribution from
regions where |f (xn )| < 1. Such regions are called contraction regions Cf of
the map, defined as
Cf = {x|f (x)| < 1}. (13.12)
1.0 1.0
0.5
−0.5
(a) (b)
0.0 −1.0
0.0 0.5 1.0 −2.0 −1.0 0.0 1.0 2.0
x x
0.8 0.8
0.6 0.6
0.4 0.4
λ
0.20 0.20
(e) (f)
0.00 0.00
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
noise intensity D noise intensity D
Fig. 13.3. Contraction regions and synchrony in chaotic maps. Left panel (a,c,e):
Logistic map f (x) = 4x(1 − x). Right panel (b,d,f ): A chaotic neuron model f (x) =
tanh(A1 x) − B tanh(A2 x) with A1 = 20, A2 = 2, and B = 1.5. The contraction
regions where |f (x)| < 1 are highlighted by thick solid lines in (a) and (b). (c) and
(d) give the LE λ, and (e) and (f ) the synchronization error E as a function of the
noise intensity D
contraction regions are intrinsic in the map f (x), but independent of the
forcing Dξ. A nonzero-mean value of the forcing D ξ , however, may shift the
new dynamical system xn+1 = f (xn ) + D ξ to different dynamical regimes,
e.g., to a periodic orbit from the previously chaotic regime.
Figure 13.3 depicts two maps which have quite different structures of con-
traction regions Cf . The well-known logistic map f (x) = 4x(1 − x) has only
a small contraction region around x = 0.5, which is manifested by the thick
line in Fig. 13.3a. The other chaotic map f (x) = tanh(A1 x) − B tanh(A2 x)
in Fig. 13.3b is a chaotic neuron model [459], which has broad contraction
regions at large |x|. In fact, these two maps have very different synchrony
properties in the presence of noise ξ uniformly distributed in [−1, 1]. As seen
in Fig. 13.3c, λ is always positive for the logistic map so that a common unbi-
ased noise cannot synchronize two identical logistic maps. The time average
of the synchrony error E = |xn − yn | displays a similar pattern (Fig. 13.3e),
but it is well above zero. Note that around D = 0.5, both λ and E reach a
minimum. Here noise makes the trajectory spend a lot of time in the con-
traction region, so that the synchrony error can reach rather small values in
some time intervals. In the chaotic neuron model, λ undergoes a transition to
negative values at D = 0.87, as shown in Fig. 13.3d. Accordingly, beyond the
276 13 Noise-Induced Synchronization
threshold, the synchrony error E decays to zero after a transient time. A more
detailed study has shown that, in the presence of strong enough noise, the sto-
chastic trajectories shift to spend a lot of time in the contraction regions [455]
where they converge to each other. When the convergence becomes dominant,
λ becomes negative, and the local instability in the regions |f (x)| > 1 cannot
keep the map chaotic. Due to the coexistence of contraction and expansion
in the phase space, close to the transition point, the separation of the trajec-
tories exhibits a very intermittent behavior [447], and the probability density
of the separation satisfies a scaling law [448]. This intermittent behavior has
been shown [460] to be on–off intermittency [461].
where (x1 , x2 ∈ RM ), the small initial difference δx(t) = x2 (t) − x1 (t) evolves
according to the linearized dynamics
where Jf (x) stands for the Jacobian matrix. This linear equation is the same
as the noise-free case (D = 0) where the system is chaotic, i.e., the maximal LE
1 T |δx(t)|
λ1 = lim ln (13.16)
T →∞ T 0 |δx(0)|
is positive. In the presence of noise, however, the trajectory x is different from
that in the noise-free system, and may explore some regions in the phase space
which are not reachable by the original chaotic system. As in the maps, for CS
to occur, λ1 should become negative, which is possible only when there exist
well-exposed contraction regions. In general, contraction regions are defined as
the region in the phase space where nearby trajectories converge. A necessary
condition for the contraction is
Cf = {x Re[Λi (x)] < 0 (i = 1, . . . , M )}, (13.17)
namely, all the M eigenvalues Λi (x) of the Jacobian matrix Jf (x) have a
negative real part.
There is different synchronization-like behavior [457] in the Rössler systems
0.075 1
0.000 0
λ1
(a) (b)
−0.075 −1
10 10
5 5
E
(c) (d)
0 0
0.0 1.0 2.0 3.0 4.0 0 10 20 30 40
noise intensity D noise intensity D
Fig. 13.4. The largest LE λ1 of the Rössler system (a) and the Lorenz system
(b). The average synchronization error E of the Rössler system (c) and the Lorenz
system (d)
80
40
z
now explore a larger region of the phase space with increasing noise intensity
D. In the CS regime D > 33.3, the trajectory is much more complex than a
smeared periodic orbit and is quite different from the external noise.
In the Lorenz system with the parameters considered here, the origin
(0, 0, 0) is a saddle point “embedded” in the chaotic attractor (i.e., the chaotic
trajectory goes into the vicinity of this point). Hence the chaotic trajectories
leaving the neighborhood of this saddle point will come back to its neigh-
borhood. Due to this homoclinic return of chaotic orbits, there exists a large
contraction region close to the stable manifold of the saddle point (Fig. 13.5).
In the presence of noise, the trajectories cannot come as close to the saddle
point as in the noise-free system; instead they explore deep into the con-
traction region. The modification of the attractor in the phase space changes
the competition between contraction and expansion. When the contraction
dominates over the expansion, the largest LE λ1 becomes negative and CS
occurs.
In the Rössler system, the trajectories spiral outward following the
guidance of the two-dimensional unstable manifold of the focus and are
fold back by the nonlinearity. A contraction region with all three Re(Λi ) < 0
does exist, but the contraction is very weak because the largest Re(Λi ) is
close to zero. In addition, in the presence of noise the system still spends
only a small portion of time in the contraction region. The contraction is not
sufficient to induce CS. There are also regions in the phase space where all
Re(Λi ) > 0, and strong enough noise (here D > 4) makes the system access
to such regions and breaks the system down easily.
To summarize, the controversy on the mechanism of noise-induced CS
of chaotic systems can be resolved in the following manner. A significant
contraction region plays a decisive role, as in the one-dimensional chaotic
maps. Noise may change the balance between contraction and expansion, and
a synchronous regime occurs when contraction becomes dominant. Whether
the noise is biased or unbiased is not the key point. If there does not exist a
contraction region, CS cannot occur with any additive common driving signal.
This understanding from the Lorenz system is very instructive to observe
noise-induced synchronization in other systems. In general, it is expected that,
in a system possessing the structure of a saddle point embedded in a chaotic
attractor and homoclinic return of orbits (homoclinic chaos), the generic exis-
tence of a large enough contraction region and the sensitivity of the system
to noisy perturbation near the saddle point will result in noise-induced syn-
chrony. Note that such a structure is typical for the spiking behavior of many
neural, chemical, and laser systems, and noise-induced synchrony may play
an important functional role there. We present two examples in the following.
The first spiking system is the Hodgkin–Huxley model [462] of thermally sen-
sitive neurons which was proposed in [463, 464] to mimic spike train patterns
13.2 Noise-Induced CS and PS in Uncoupled Chaotic Oscillators 279
observed in electroreceptors from dogfish and catfish, and from facial cold
receptors and hypothalamic neurons of the rat. The model equations read
dV
CM = −Il − Id − Ir − Isd − Isr + Dξ(t)
dt
ar φ(T )(ar∞ − ar )
= ,
dt τr
asd φ(T )(asd∞ − asd )
= ,
dt τsd
asr φ(T )(−ηIsd − θasr )
= , (13.20)
dt τsr
with Id = ρ(T )gd ad∞ (V − Vd ) and Ik = ρ(T )gk ak (V − Vk ), (k = r, sd), where
(T −T0 )/10 (T −T0 )/10
ak∞ = [1+exp(−sk (V −V0k ))]−1 and ρ(T ) = A1 , φ(T ) = A2 ,
ξ(t) is the Gaussian noise with intensity D. Here V is the membrane potential,
Il is the leakage current, and Id and Ir are fast currents representing Na and
K channels (see also Sect. 2.7.1).
In the chaotic spiking regime, the system possesses an unstable steady state
which has both stable and unstable two-dimensional manifolds. Therefore, it
is a saddle point S. A trajectory starting close to the stable manifold will
approach the neighborhood of S and leave it along the unstable directions.
This saddle point S is embedded in the chaotic attractor, i.e., a typical chaotic
trajectory may have very close recurrence to S after a sequence of spikes. The
eigenvalues corresponding to the stable manifold are real (λ3 < 0, λ4 < 0), but
the eigenvalues corresponding to the unstable manifold are complex (λ1,2 =
μ ± iν) and −λ3,4 > μ [465], and the chaotic dynamics resulting from this
Shilnikov condition [466] is a typical mechanism of chaotic spiking in neuron
models [467].
Close to the stable manifold, a large contraction region exists in the phase
space where nearby trajectories converge, while close to the unstable mani-
fold nearby trajectories diverge. The contraction and expansion in this high-
dimensional system can be manifested, by the largest local LE λτ ,
1 |δx(t)|
λτ (t) = ln , (13.21)
τ |δx(t − τ )|
where |δx(t)| is a small distance between two trajectories in the phase space at
time t, and τ is a finite time. λτ measures the average expansion or contraction
rate during the finite interval τ . Here τ = 50 ms is about half of the duration
of a single spike, but is much smaller than the average interspike interval
T ∼ 400 ms. With this τ value, quick and large changes of local stability
along the orbits during a spike have been smoothed out considerably and we
can see clearly the changes of the stability close to the saddle point where the
orbits slow down.
A typical chaotic spike sequence of the neuron at T = 10 (without noise
input) is shown in Fig. 13.6a, along with the largest local LE λτ in Fig. 13.6b.
280 13 Noise-Induced Synchronization
(a)
0
V (mv)
S
− 40
− 80
(b) 120
Λτ (1/s)
0
S
− 120
2.0 3.0 4.0 5.0 6.0
time (s)
Fig. 13.6. Spike train (a) and local LE (13.21) (b) of the chaotic neuron (13.20)
at T = 10, without noise D = 0. The arrows indicate the saddle point S
The saddle point S is indicated by arrows in Fig. 13.6 close to t = 2.5 and
t = 4.5. The neuron generates a sequence of spikes between successive returns
to the neighborhood of S. After a a few spikes away from S in the phase
space, the orbit approaches S, and is guided by the stable manifold as seen
by negative λτ ; then it departs from S following the guidance of the unstable
manifold for a long time, as manifested by small positive λτ . The average
contraction rate is stronger than the expansion rate since the eigenvalues
satisfy −Λ > μ. A single spike (e.g., close to t = 2.8 and t = 5.3 in Fig. 13.6)
follows this recurrence to S. Such a single spike is well separated from the
next burst, because the orbit cycles S and slows down around S. For each
spike, λτ is positive during the activation phase, while it is negative during
the relaxation phase. In the absence of noise, the spike sequence is chaotic so
that the largest LE is positive, i.e., λ1 = λτ > 0.
Note that, in this system, the dynamics is sensitive to the noisy input close
to the saddle point, while the spikes are not affected much. This property is of
importance for biological information processing using spike trains which are
well defined even in the presence of a fluctuating input. The synchronization
behavior in this system is determined by the competition between contraction
and expansion in the phase space. Since the contraction rate is larger than
the expansion rate around S, in the presence of a suitable level of noise, the
orbits can still approach S and experience contraction, while they cannot
follow the guidance of the weakly unstable manifold for a long time. There
are two consequences of this behavior:
1. Those long intervals in the noise-free case resulting from following the
unstable directions of S have been reduced considerably.
2. The expansion degree is reduced correspondingly compared to the noise-
free case. As a result, the largest LE λ1 = λτ becomes negative at
a threshold noise level Dc ≈ 3.5 pA cm−2 . Beyond the threshold, CS is
achieved for two identical neurons forced by a common Gaussian noise, as
13.2 Noise-Induced CS and PS in Uncoupled Chaotic Oscillators 281
2
(a)
λ1(1/s)
1
1
1.0
(b)
E
0.5
0.0
0.0 1.0 2.0 3.0 4.0 5.0
Very similar noise-induced CS of spike trains due to a saddle point S has been
observed in an experiment of homoclinic chaotic lasers [469].
The experimental setup consists of a CO2 laser with an intracavity loss
modulator (parts 1–4 in Fig. 13.8), driven by a feedback signal which is pro-
portional to the laser output intensity (parts 6–8 in Fig. 13.8). This system
is operating in a homoclinic chaotic regime where the laser output consists
of a chaotic sequence of spikes. To investigate the role of external noise, a
Gaussian noise generator with a frequency cut-off at 50 kHz is inserted into
the feedback loop.
Before presenting the results of the experiments, we analyze the model of
this laser. The dynamical equations describing appropriately the experimental
system read:
To digital scope
1 1
4
2 6
3 3 7
R
9 8
B0
Noise source
(b) S
(a)
T
x1
Fig. 13.9. (a) Laser output intensity x1 of the noise-free model (13.22).
(b) Projection of the trajectory in the 3D phase space (x1 , x2 , x6 )
weak expansion along the unstable manifold, which can be described approx-
imately as
X(t) ∼ X0 exp[Λu (t − t0 )] cos ω(t − t0 ), (13.23)
where Λu ± iω are the leading eigenvalues of the unstable manifold of S,
and X0 is the distance from S at any reinjection time t0 . Thus, the smaller
X0 is, the longer is the time taken to spiral out. In the presence of noise,
the trajectory on average cannot come closer to S than the noise level, and
perform those oscillations very close to S, while the dynamics along the stable
manifold is not much affected by the noise due to the strong contraction.
With a larger X0 , the system spends a shorter time following the guidance
of the unstable manifold, resulting in a reduced degree of expansion. This
highly nonuniform sensitivity to noise perturbation changes the competition
between contraction and expansion, and contraction may become dominant
at large enough D. The largest LE λ1 in the model changes as a function
of the noise intensity D. As seen in Fig. 13.10a, λ1 undergoes a transition
from a positive to a negative value at Dc ≈ 0.0031. Beyond Dc , two identical
laser models x and y with different initial conditions but with the same noisy
driving Dξ(t) achieve CS regime after a transient, quantified by the vanishing
|x1 −y1 |
normalized synchrony error E = |x 1 −x1 |
. At larger noise intensities, the
expansion becomes again significant, and CS is lost when λ1 becomes positive
for D > 0.052. Notice that, even when λ1 < 0, the trajectories still have access
to the expansion region where small distances between them grow temporally.
As a result, when the systems are subjected to additional perturbations, CS is
lost intermittently, especially for D close to the critical values. Actually, in the
experimental laser system, there always exists a small intrinsic noise source.
To take into account this intrinsic noise into the model, we introduce into
the equations x6 an equivalent amount of independent noise (with amplitude
D1 = 0.0005) in addition to the common one Dξ(t). By comparison, it is
evident that the sharp transition to CS in fully identical model systems is
smeared out.
284 13 Noise-Induced Synchronization
10 1.7 1.7
E E
λ1 (1/ms)
0 0.0 0.0
(a) (b)
10 −1.7 −1.7
0.00 0.02 0.04 0.06 0 50 100 150
noise intensity D noise intensity D (mV)
Fig. 13.10. Noise-induced CS in the model (13.22) (a) and the experiment (b)
(Fig. 13.8). (a) Dotted line: the largest LE λ1 ; solid line: normalized error of syn-
chrony E between two fully identical laser models x and y; diamonds: E between
two lasers with a small independent noise (intensity D1 = 0.0005) equivalent to the
intrinsic noise in the experiments
Fig. 13.11. Time series of output intensities (arbitrary units) of two lasers x1
(solid lines) and y1 (dotted lines) with a common noise. Left panel (a–c): model
systems (13.22) including independent noise (amplitude D1 = 0.0005 to intrinsic
noise level). Right panel (d–f ): experimental system (Fig. 13.8)
In Sect. 13.2.1, we have shown that under certain conditions noise may induce
CS in chaotic systems, depending on the structure of the phase space. It is
interesting to study whether noise can also induce weaker degrees of synchro-
nization, similar to temporal locking of phases as in PS, when it is not strong
enough or not able to induce CS.
In Chap. 4, we have shown that in many systems there is the regime of PS
associated with the transition of one of zero LE to negative values. In time-
continuous autonomous chaotic systems ẋ = f (x) (f : Rn → Rn ), there is a
zero LE λ0 corresponding to perturbations of the motion along the trajectory,
which, in phase-coherent chaotic systems, can be connected to the phase. In a
system subjected to a weak noise, ẋ = f (x)+ξ, most time it is |f (x)| |ξ| and
one can also roughly speak about a motion along the trajectory and connect
the original zero LE to it and link it to the phase dynamics. Numerical results
286 13 Noise-Induced Synchronization
show that λ0 becomes negative for strong enough noise, and in parallel, the
phases of the two slightly nonidentical systems coupled only by common noise
become statistically correlated [457].
To study PS due to noise, two systems with a small parameter mismatch
are considered: particularly, ω1 = 0.97 and ω2 = 0.99 in the Rössler sys-
tem (13.18) and σ1 = 10, ρ1 = 28 and σ2 = 10.2, ρ2 = 28.5 in the Lorenz
system (13.19). This slight parameter difference does not change the LE spec-
tra of the systems much.
In noisy chaotic systems, the phase of the dynamics is no longer coherent,
i.e., a linearly increasing function of time. A phase linked to λ0 now cannot
be rigorously defined as in the noise-free case. Nevertheless, as discussed in
Chaps. 2–4, we can practically calculate a phase variable as in the determin-
istic system by the method of Poincaré section, e.g.,
t − τk
φ(t) = 2π[k + ], τk < t < τk+1 , (13.24)
τk+1 − τk
where τk and τk+1 are two successive crossings of a Poincaré section after
cycling a reference point (unstable fixed point of the noise-free system). In
stochastic systems, it is impossible to observe perfect synchrony of phases
φ1 and φ2 , i.e., |φ1 − φ2 | < const. An appropriate approach to study PS
in stochastic systems is to regard the phases in a stochastic manner, i.e., to
compute the distribution of the cyclic phase difference, P (Δφ), on [−π, π] [473,
474]. A peak in P (Δφ) manifests a preferred phase difference between the
systems. Preferred phase differences are expected to occur at least when λ0
becomes appreciably negative.
Without noise the two nonidentical Rössler systems are not phase synchro-
nized and the phase difference decreases almost linearly (Fig. 13.12a, D = 0).
Accordingly, the distribution of the phase differences Δφ on [−π, π] is almost
uniform (Fig. 13.12b). While for strong enough noise, where λ0 < 0, one
observes many plateaus in the phase difference, i.e., many phase locking epochs
(Fig. 13.12a, D = 3.0), and this is manifested by a pronounced peak around
Δφ = 0 in the distribution P (Δφ) (Fig. 13.12d); this demonstrates a noise-
induced imperfect PS. For D close to the transition of λ0 , the peak is not as
pronounced and is not located around Δφ = 0 (Fig. 13.12c). Similar properties
are also observed in the Lorenz system.
The degree of noise-induced imperfect PS can be measured quantitatively
by the mutual information between the cyclic phase dynamics on [−π, π] of
the two systems
p(i, j)
M1 = p(i, j) ln , (13.25)
i,j
p 1 (i)p2 (j)
where p1 (i) and p2 (j) are the probabilities when the phases φ1 and φ2 are in
the ith and jth bins, respectively, and p(i, j) is the joint probability that φ1
13.2 Noise-Induced CS and PS in Uncoupled Chaotic Oscillators 287
−70
(a) D=3
−100
Δφ
D=0
−130
5000 5500 6000
time
0.01
0.00−π π −π π −π π
0 0 0
Δφ Δφ Δφ
Fig. 13.12. (a) Time series of the phase difference Δφ of two nonidentical Rössler
systems (ω1 = 0.97 and ω2 = 0.99), at noise intensity D = 0 and D = 3. In this
presentation, the phases are unwrapped. The distribution of the phase difference is
modulated into [−π, π] for D = 0 (b), D = 0.7 (c), and D = 3 (d)
1
0.00 0
(a)
λ0
−1 (b)
−0.05 −2
0
−2
10−1
10 10−2
−3 10−3
M
10 (d)
10
−4 (c) 10−4
−5 10 −5
10 10 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 5 10 15 20
noise intensity D noise intensity D
Fig. 13.13. The second LE λ0 of the Rössler system (a) and the Lorenz system
(b). Normalized mutual information M of the Rössler system (c) and the Lorenz
system (d)
is in the ith bin and φ2 in the jth bin. The number of bins of [−π, π] in our
simulations is N = 100. M1 is normalized into [0, 1] as M = M1 /Sm , where
Sm = ln N is the Shannon entropy of the uniform distribution of p1 and p2 .
Numerical results for M are shown in Fig. 13.13 along with the second LE λ0
as a function of the noise intensity D. Due to the incoherence of the phases,
an exact correspondence between the transition of λ0 and phase synchrony
would not be expected. Nevertheless, when λ0 becomes appreciably negative,
M increases rapidly, indicating an increasing degree of PS.
288 13 Noise-Induced Synchronization
0.000
−π 0 π −π 0 π −π 0 π
θ θ θ
0.2
(a) (c) (e)
P(T)
0.1
0.0
0.2
(b) (d) (f)
P(T)
0.1
0.0
0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0
T (ms) T (ms) T (ms)
100
D=0
70
θ(t) 40 D = 0.01
10
D = 0.0005
−20
0 100 200 300 400 500
time (ms)
Fig. 13.16. Phase difference between the laser model and the driving signal at
various noise intensities. The forcing amplitude is A = 0.01, and the signal period
Te is the same as in Fig. 13.15b,d,f for D = 0, D = 0.0005, and D = 0.01, respectively
noise eliminates most of the oscillations around S (see Fig. 13.11b for a similar
behavior); the fine structure of the peaks is smeared out and P (T ) becomes a
unimodal peak with a lower height (Fig. 13.15e). Note that the average value
T0 (D) of T decreases with D. Consequently, the average spiking rate of the
unforced laser model, f0 (D) = 1/T0 (D), is an increasing function of D.
As a result of these noise-induced changes in the time scales, the model
displays a quite different response to a weak signal (A = 0.01) with a frequency
fe = f0 (D) = 1/T0 (D), i.e., equal to the average spiking rate of the unforced
model. At D = 0, P (T ) of the forced model still has many peaks (Fig. 13.15b),
while at D = 0.0005, T is sharply distributed around the signal period Te =
T0 (D) (Fig. 13.15d). However, at larger intensity D = 0.01, P (T ) becomes
lower and broader again (Fig. 13.15f). To examine phase synchronization due
to the driving signal, the phase difference θ(t) = φ(t) − 2πfe t is computed.
Here the phase φ(t) of the laser spike sequence is simply defined by (13.24),
where τk now is the spiking time of the kth spike. Figure 13.16 displays the
phase difference θ corresponding to Fig. 13.15b,d,f. At D = 0, the phase of the
laser model is not locked by the external forcing. On the contrary, with a small
noise D = 0.0005, phase slips occur very rarely and phase locking becomes
almost perfect when noise generates a pronounced time scale in the system. At
stronger intensity D = 0.01, noise becomes dominant over the signal around
the saddle S, and it induces many random-like phase slips. The behavior is
similar for driving frequencies close to f0 (D).
Now we study the synchronization region (1:1 response) of the laser model
systematically in the parameter space of the driving amplitude A and the
relative initial frequency difference Δω = (fe − f0 (D))/f0 (D). The actual
relative frequency difference in the presence of the signal is calculated as
ΔΩ = (f − fe )/f0 (D), where f = 1/ T t is the average spiking rate of the
model subject to the pumping modulations. The results are shown in Fig. 13.17
for three different noise intensities. In the noise-free model (Fig. 13.17a), the
13.3 Noise-Enhanced PS in Weakly Coupled Chaotic Oscillators 291
0.015
0.010
A
0.005
(a) (b) (c)
0.000
−0.3 0.0 0.3 −0.3 0.0 0.3 −0.3 0.0 0.3
Δω Δω Δω
Fig. 13.17. Synchronization region of the laser model at various noise intensities.
A dot is plotted when |ΔΩ| ≤ 0.003: (a) D = 0, (b) D = 0.0005, and (c) D = 0.005
|T / T − 1| < 1. (13.30)
the spiking sequence of the laser, the response of the phase to perturbations
by the signal is strongly correlated with the amplitude A. In fact, in Chap. 3,
it has been shown that PS of the Lorenz system to periodic signal is imperfect:
there is no region as in the Arnold tongue of phase-coherent oscillators where
the phase is locked forever [477].
Our analysis of the laser system has shown that, when the system is kicked
away from the neighborhood of the fixed point by noise, the degree of phase-
coherence is increased so that the system shows an enhanced synchronization
to weak periodic signal. The locking tongues are now similar to phase-coherent
oscillators in the presence of noise. Such a property seems to be general also in
excitable systems where noise kicks the system to escape the stable fixed point
to generate a sustained oscillation. Noise-enhanced synchrony of excitable
systems will be shown in Sect. 13.4.
Here ω1 = 0.97 and ω2 = 0.99. The added noise Dξi (t) (i = 1, 2) is now
composed of a common part e(t) and an independent part ηi (t), satisfying
√ √
ξi (t) = Re(t) + 1 − Rηi (t). (13.32)
Both e(t) and ηi (t) are assumed to be Gaussian noise and δ-correlated in time.
This description allows independent variations of the correlation R between
ξ1 and ξ2 and the noise intensity D.
Without noise, the two oscillators achieve PS for the coupling strength
ε > εps = 0.0208, where εps is the transition point to PS. In the weak cou-
pling region ε < εps , the phases are not yet fully synchronized, while when ε
approaches εps , appreciable PS epochs can be observed between phase slips.
For example, at ε = 0.0205, the phase difference Δφ = φ1 − φ2 decreases con-
tinuously. But there are many epochs of PS interrupted by phase slips; and
typically the PS epochs last for about 300 cycles of oscillations (Fig. 13.18).
Adding a proper amount of noise to the two oscillators (e.g., D = 0.1) can
prolongate remarkably the duration of the synchronization epochs: the two
oscillators maintain phase synchronization for a period of about 3,000 cycles
of oscillations. However, for stronger noise (e.g., D = 0.3), phase slips occur
more frequently again.
This phenomenon of noise-enhanced phase synchronization can be quanti-
fied by the duration τ of the PS epochs. It is found that the average duration
13.3 Noise-Enhanced PS in Weakly Coupled Chaotic Oscillators 293
104
0 a b
D=0.1
50
103
<τ>
Δφ
D=0.3
ε=0.0205
100 ε=0.0200
ε=0.0195
D=0 102
ε=0.0190
150
0 25000 50000 0.0 0.1 0.2 0.3 0.4
time noise intensity D
0 0
(a) (d)
Δφ
Δφ
−20 −20
9 6.4
(b) (e)
6 6.3
T4/4
ΔX4
3 6.2
0 6.1
6 6.4
4 (c) (f)
6.3
ΔX2
T4/4
2 6.2
0 6.1
2000 4000 6000 2000 4000 6000
t t
Fig. 13.19. Illustration of phase slips (a, d) induced by unlocked UPOs (b, e,
period-4 for oscillator 1; c, f, period-2 for oscillator 2) at ε = 0.0205. The insets
in (b) and (c) show the unlocked UPOs around t = 2, 500. T4 /4 in (e) and (f ) is
the average return time calculated by every four returns to the Poincaré section in
system (13.31)
trajectory is very close to it. It is seen that phase slips occur between a period-
4 UPO in oscillator 1 and a period-2 UPO in oscillator 2 which are followed
closely by the systems for a fairly long time (∼ 30 cycles). While most orbits
are locked with return times fluctuating around a common value (T = 6.24),
these UPOs have clearly much smaller and larger return times (Fig. 13.19e,f),
thus remain unlocked by the coupling. With a noise of D = 0.1, such a long
time staying close to UPOs is rarely observed, and meanwhile most of the
phase slips are eliminated (Fig. 13.18). At stronger noise, e.g., D = 0.3, phase
slips develop quickly when the oscillators come to some orbits with quite large
differences in the return times, which cannot follow UPOs closely.
Noise-enhanced PS thus can be explained as follows. Noise exerts two
effects (1) it prevents the system from staying close to the unlocked UPOs
for long enough times to allow a phase slip to occur and (2) it generates fluc-
tuations in the return times and may induce phase slips of locked orbits, as
it does in coupled periodic orbits. The degree of PS is enhanced when (1)
dominates over (2) at a weak noise level, while it is degraded again when (2)
becomes dominant at large noise. There thus exists an optimal noise inten-
sity, yielding the maximal enhancement as a result of the competition between
these two influences and it is a resonance-like phenomenon. At a smaller cou-
pling strength ε, more orbits become unlocked, and phase slips may develop
already during a shorter time τsl when the oscillators approach some unlocked
orbits. When noise prevents a phase slip, the trajectories may approach other
13.3 Noise-Enhanced PS in Weakly Coupled Chaotic Oscillators 295
1.06 20
5 (a)
1.05
Δφ1,5
40
1.04 4 (b)
60
Ωi
1.03
3
1.02 4 (c)
ΔX1
2
1.01 2
1
1.00 0
0.00 0.05 0.10 2000 3000 4000 5000
ε time
staying close to unlocked UPOs, it also induces phase slips among clustered
oscillators. As a whole, there is no appreciable enhancement of PS of the
ensemble by independent noise. In the following, we consider a global noise
R = 1, which is common to all the oscillators in the ensemble.
In the globally coupled ensemble, the transition to PS can be quantitatively
described by the mean order parameter r(t) , where
r(t) = | Pj (t)|/ |Pj (t)| (13.34)
j j
quantifies the degree of clustering of the state vectors Pj (t) = (xj , yj ) in the
phase space at time instant t. This definition of the order parameter based
on state points in the phase space has the same meaning as that defined in
(5.43) based on the phase of the oscillators. For unsynchronized oscillators,
the state vectors are scattered in the phase space and r is close to zero; while
r > 0 indicates a degree of clustering of the state vectors. Due to clustering,
a collective oscillation emerges in the ensemble of the oscillators. So the order
parameter r(t) is closely related to the oscillation amplitude of the mean field
X, which is quantified by the variance var(X) of X over time. The frequency
locking behavior can be quantified by σΩ which is defined as the standard
deviation of the mean frequency Ωj = φ˙j t of the oscillators in the ensemble.
In the absence of noise, all the oscillators are locked to a common frequency
and σΩ becomes zero at large enough coupling ε. However, a smaller σΩ may
not always indicate a stronger degree of synchronization, especially in the
frequency clustering region where σΩ may not be small but the degree of
synchronization is actually high.
As illustrated in Fig. 13.21, in the noise-free ensemble consisting of 1,000
elements, all the measures vary slowly until ε ≈ 0.043 where they start to
change quickly, and finally global PS is achieved at εps ≈ 0.093. In the
crossover region 0.043 < ε < 0.093, many oscillators are locked into clus-
ters with a common frequency within a cluster and different frequencies for
different clusters. A collective oscillation emerges when a significantly large
cluster is formed, and its amplitudes increase with ε when more oscillators are
included into the leading cluster. Unlike in the case of two coupled oscillators,
when independent noise (R = 0) is added to the ensemble, the PS regime is
degraded, as indicated by smaller var(X) and r , because noise degrades the
synchronization regime of those locked oscillators significantly. In the cluster-
ing region, σΩ becomes smaller than that of the noise-free ensemble, which,
however, does not mean a higher degree of PS. In contrast, a global noise
(R = 1) which is common to all the oscillators can enhance PS significantly
in the weak coupling regime: σΩ is considerably smaller, and var(X) and r
are clearly larger than those in the absence of noise.
For a fixed coupling strength ε < εps , there is an optimal noise inten-
sity D which enhances synchronization the most, and the mean field displays
the most coherent oscillation. For smaller D, the effective coupling due to
the global noise does not enhance synchronization, and for a too large noise
13.3 Noise-Enhanced PS in Weakly Coupled Chaotic Oscillators 297
50 D=0
40 D=0.7, R=0
var(X)
30 D=0.7, R=1
20
10 (a)
0
0.02
σΩ
0.01
(b)
0.00
1.0
<r>
0.5
(c)
0.0
0.00 0.05 0.10 0.15
ε
Fig. 13.21. Transition to PS in an ensemble of N = 1, 000 globally coupled chaotic
Rössler oscillators with or without noise. (a) The variance of the mean field X.
(b) Standard deviation of the distribution of the mean frequency Ωj . (c) The order
parameter (13.34)
Two Oscillators
We start with two coupled oscillators. The two electrodes have a small fre-
quency mismatch (Δω = 14 mHZ) due to surface heterogeneity. With an
added weak coupling, a PS regime occurs at about εps = 0.08; i.e., the fre-
quencies of both oscillators become equal.
Similar to two coupled Rössler oscillators in Sect. 13.3.2, just before the
threshold, one can observe clearly even in this experiment the correspondence
between phase slips and UPOs. For example, at ε = 0.06, the observed fre-
quency mismatch (ΔΩ = 5 mHz) is smaller than that seen without coupling,
however, the coupling is not strong enough for PS. During the time of the
experiment, one phase slip was observed (Fig. 13.24a). The analysis of the time
series of the two oscillators shows that the phase slip occurs when both oscilla-
tors approach the neighborhood of an unlocked period-3 UPO. In Fig. 13.24b,
13.3 Noise-Enhanced PS in Weakly Coupled Chaotic Oscillators 299
Counter Electrode
Reference Electrode
V
Working Electrode
Potentiostat
Rind
Rcoll
Fig. 13.23. Schematic diagram of the experimental setup: globally coupled elec-
trochemical oscillators
3 (a)
0
Δφ
3
6
0.10 (b)
ΔX3/mA
0.05
0.00
0 50 100 150
t /s
Fig. 13.24. Two coupled (ε = 0.06) chaotic electrochemical oscillators just below
εP S = 0.08. (a) Phase difference between both oscillators vs. time. (b) The difference
between the next return values of the current maxima (ΔX3 = |Xn − Xn−3 |, where
Xn is the nth maximum) of the two oscillators (solid : oscillator 1, dashed : oscillator
2) as a function of time
the ΔX3 values are shown as a function of time for both oscillators. The coin-
cidence of the approach of the unlocked UPOs and the phase slip confirms the
numerical predictions about the dynamics close to but below εps .
300 13 Noise-Induced Synchronization
10
15 (a) (d)
10 0
Δφ
5
0 −10
0.10 (b) 0.10 (e)
ΔX4/mA
0.05 0.05
0.00 0.00
0.10 (c) 0.10 (f)
ΔX4/mA
0.05 0.05
0.00 0.00
0 50 100 150 0 50 100 150
t /s t /s
Fig. 13.25. Two coupled (ε = 0.04) chaotic oscillators without (left panel, a–c)
and with (right panel, d–f ) small amounts of common, zero-mean Gaussian noise
(standard deviation of 3 × 10−4 V, measured at 200 Hz). Top row : phase difference
between the oscillators vs. time; middle and bottom row : the difference between the
next return values of the maxima of the oscillators, ΔX4 = |Xn − Xn−4 | (middle:
oscillator 1, bottom: oscillator 2), vs. time. Rtot = 500 Ω, V0 = 1.350 V
Array of 64 Oscillators
Fig. 13.27. Dynamics of 64 electrodes as in Fig. 13.26, but with optimal noise
intensity, D = 4 mV
Fig. 13.28. Dynamics of 64 electrodes as in Fig. 13.26, but with excessive noise,
D = 10 mV
Ωp
β = Ap , (13.36)
ΔΩ
where Ωp is the frequency of the main peak in the spectrum of h(t), Ap is
the peak height mainly depending on the amplitude of h(t), and ΔΩ is the
half width of the peak reflecting temporal randomness of h(t). Defined in this
304 13 Noise-Induced Synchronization
way, β measures the sharpness of the main peak in the spectra. A large value
of β indicates that the oscillations of h(t) are rather periodic, thus are highly
coherent.
The results of these measures are shown as a function of the noise inten-
sity in Fig. 13.29 for ε = 0 and ε = 0.014. The mean field increases in a
monotone way for ε = 0, showing that noise can induce a degree of PS even
for uncoupled oscillators, as is consistent with the prediction in Sect. 13.2.2.
With the weak coupling ε = 0.014, the mean field shows a maximum, indicat-
ing a resonant-like behavior, which can also be observed in ρ and β – there is
a small variation for ε = 0 and a resonance curve for ε = 0.014. However, the
frequency distribution (Fig. 13.29c) does not clearly exhibit the trend above:
although for ε = 0.014 there is a minimum at the optimal noise intensity, for
ε = 0 there are larger variations.
The experiments on ensembles of chaotic oscillators confirm the noise-
induced effects predicted by numerical calculations. At an optimal common
a b
−2
0.14
−2.5 0.12
log10 var(h)
0.1
−3
ρ
0.08
0.06
−3.5
0.04
−4 0.02
0 2 4 6 8 10 0 2 4 6 8 10
x 10−3 c d
7000
12 ε=0
6000 ε=0.014
10 5000
σω /ωmean
4000
β
8
3000
6 2000
1000
4
0 2 4 6 8 10 0 2 4 6 8 10
Noise intensity /mV
Fig. 13.29. Different measures of the collective behavior as a function of noise
intensity for ε = 0 (squares) and ε = 0.014 (circles). (a) Variance of mean field
var(h). (b) Temporal PS index ρ. (c) Relative standard deviation of frequencies
σω /ωmean . (d) Coherence factor β
13.4 Noise-Enhanced Synchronization-Like Phenomena 305
away from the steady state. To make the study more general, we suppose
that ai is not identical for the elements, but is uniformly distributed ai ∈
(−1.1, −1). This implementation of the model is important because many
real systems are diffusively coupled, and nonidentity is also more natural in
physics, engineering, or biology. With nonidentical ai , the uncoupled elements
will have different average firing frequencies in the presence of the same level
of noise ξi (t). We assume that the Gaussian noise is uncorrelated in different
elements, i.e., ξi (t)ξj (t ) = δij δ(t − t ). Periodic boundary conditions are
considered. The parameter of the time scale is fixed at = 0.01.
To characterize synchronous behavior in the chain of these nonidentical
excitable systems, we introduce the phase of the elements as in (13.24), where
τk is the time of the kth firing of the element defined by threshold crossing of
xj (t) at x = 1.0. The quantity
φi − φi+1
ρi = sin2 (13.38)
2
20
10
S
−0.5
0 −1
0.5
−0.5 −1.5
−1.5
−2.5 −2 lg(D)
lg(d)
Fig. 13.30. Signal-to-noise ratio S in the parameter space (log10 (d), log10 (D)) of
a coupled chain of nonidentical FHN neurons (13.37) with N = 100. The black dots
correspond to the patterns in Figs. 13.31–13.34
The main features of the system are illustrated in Fig. 13.30, where the
results of S in the parameter space (log10 (d), log10 (D)) are shown for a chain
of N = 100 elements. For a fixed coupling strength d, S increases first with
the noise level D, reaches a maximum, then decreases again; i.e., it shows a
typical resonance-like behavior. In general, for a stronger coupling, a higher
level of noise is needed to excite the system. Similarly, for a fixed noise level
D, S increases with increasing d until it reaches an optimal value; after that,
it decreases again.
From Fig. 13.30, one can observe several dynamical regimes in the system.
For very weak coupling (d < 10−2 ), the firing of the elements are essentially
independent, because a noise-induced firing of an element cannot excite its
neighboring lattices. A typical spatiotemporal pattern of xi and ρi for weakly
coupled elements subject to relatively weak noise is shown in Fig. 13.31. Both
the spatial and temporal behaviors are quite irregular. Due to the independent
firing, the frequencies and phases are not synchronized, and the distribution
P (T ) is broad.
With an increase of the coupling strength, the system becomes sensitive to
weak noise because the firing events induced by noise now become the source of
excitation of the neighboring elements and leads to partial synchrony, as seen
in Fig. 13.32. This mutual excitation enhances the coherence of the motion
in the coupled system, as indicated by a narrower distribution P (T ) and a
larger S. The lattice displays clusters of synchrony. The clusters break and
reunite during the evolution, so that each element has a slightly different firing
frequency.
The next regime where S takes the largest values (S ∼ 18) is the most
interesting one, because the system performs a quite regular motion globally,
308 13 Noise-Induced Synchronization
Fig. 13.31. Synchrony in a chain of coupled noisy excitable FHN neurons (13.37)
with d = 0.005 and D = 0.02. The upper panels show the spatiotemporal structure
of xi and ρi . The time step is 0.2. The lower panels show the average firing frequency
in the chain and the distribution of the pulse duration
as seen in Fig. 13.33. All elements are locked to a relatively large firing fre-
quency, and the distribution of the pulse duration becomes very sharp. After
that, with a stronger coupling, the system keeps the global synchrony regime,
however, the temporal behavior becomes irregular again (Fig. 13.34), corre-
sponding to a decreased S.
The locations of the four representative dynamical regimes are indicated
by the black dots in Fig. 13.30. These regimes are typical for different sizes of
the chain. For a larger chain, the regime of global regular motion (S large) is
wider in the parameter space [484].
Further analysis [485] has shown that high coherence of the phase synchro-
nized oscillations in chains of coupled FHN neurons is supported by spatially
independent noise components, and a correlated or global noise reduces the
temporal coherence of the spiking. In a certain regime of noise intensity and
coupling strength, the disorder in the excitability (inhomogeneous parameter
ai ) can further enhance the coherence.
13.4 Noise-Enhanced Synchronization-Like Phenomena 309
1 d
Dφ = (Δφ − Δφ )2 , (13.40)
2 dt
where Δφ is the phase difference between an element and the reference ele-
ment, and · denotes average over the elements. In the regime of optimal
noise, Dφ becomes minimal, indicating the best synchrony behavior.
A geometric approach is introduced in [487] for understanding the phe-
nomenon of PS in coupled nonlinear systems in the presence of additive noise.
Fig. 13.33. As in Fig. 13.31, but for d = 0.25 and D = 0.07. The spatiotemporal
plot of ρ is black, indicating no phase slips
x3i
ẋi = xi − − yi + A cos Ωt + d(xi+1 + xi−1 − 2xi ),
3
y˙i = xi − a + Dξi . (13.41)
Here the parameters are in the excitable regime, = 0.01 and a = −1.05,
with the coupling strength d = 0.05. To demonstrate the significant role of
the coupling, the synchronous behavior of the coupled chain (N = 30) to the
external forcing is compared with a single uncoupled neuron (N = 1).
Now a weak noise D = 10−1.75 is added. Both systems of a single (N = 1)
uncoupled neuron and an array (N = 30) of coupled FHN neurons show
a broad distribution of the spiking intervals T (Fig. 13.35a), although the
coupling has reduced slightly the probability of very long intervals. The two
systems, however, have quite different responses to the same subthreshold
signal with a period Te close to the peak of P (T ) (Fig. 13.35, vertical dashed
lines). An uncoupled neuron may fail to fire a spike at some periods of the
13.4 Noise-Enhanced Synchronization-Like Phenomena 311
Fig. 13.34. As in Fig. 13.31, but for d = 1.0 and D = 0.04. The spatiotemporal
plot of ρ is black, indicating no phase slips
−1 0
10 (a) N=30 (c)
(b)
−2
10 −100
N=1
P(T)
Δφ
−3 N=1
10 N=1
−4
−200
10 N=30 N=30
10
−5 −300
0 10 20 30 40 0 5 10 15 20 25 0 1000 2000
0.06
A
0.04
0.02
(a) (b) (c)
0.00
0.08
0.06
A
0.04
0.02
(d) (e) (f)
0.00
0.8 1.2 1.6 2.0 0.8 1.2 1.6 2.0 0.8 1.2 1.6 2.0
Ω Ω Ω
Fig. 13.36. Comparison of the locking behavior of N = 1 (upper panel ) and
N = 30 (lower panel ) neurons (13.41) at various noise intensities D = 10−1.75
(a, d), D = 10−1.5 (b, e), and D = 10−1.3 (c, f ). Filled dots: effective locking region
(|Δω| < 0.002) of the noisy systems; dashed line: the threshold beyond which the
noise-free systems generate sustained spike trains. Above the solid lines is the 1 : 1
superthreshold locking region of the noise-free systems
0.08
1:1 2:3
1:2
0.06
0.04
A
0.02
2:1 1:1 2:3
1:2
0.00
0.5 1.5 2.5 3.5 4.5
Ω
Fig. 13.37. m : n Arnold tongues for a chain ((13.41), N = 30) with D = 10−1.5
(dots). The dashed line shows the spiking threshold and the solid lines are the border
lines of the m : n superthreshold locking region at D = 0
system remains intact, while a large subthreshold locking region at Ω > ω0 (D)
emerges (Fig. 13.36), where ω0 (D) is the mean firing frequency induced by
noise of intensity D in the chain without external forcing. At D = 10−1.5 , lock-
ing can be achieved with almost vanishing A when Ω ≈ ω0 (D) (Fig. 13.36e).
At D = 10−1.3 , the locking region shrinks a bit (Fig. 13.36f), and it shrinks
further for even larger D.
Higher order m : n locking regimes, where each cell generates m spikes
during every n periods of the signal, are also observed (Fig. 13.37). It is seen
again that an m : n locking can be achieved with almost vanishing A when
mΩ ≈ nω0 (D). The locking regions are no longer confined by the border lines
of the superthreshold locking regions of the noise-free system; in contrast, they
n
become centered around m ω0 (D) which moves with D. We emphasize that an
m : 1 (m > 1) superthreshold locking region does not exist in the noise-free
system, while in the noisy array, a 2 : 1 region is observable.
These results show that the interplay between coupling and noise can have
a significant role in enhancing the resonant response of excitable systems due
to external forcing. Resonances and locking occur due to a matching between
the noise-controlled time scale and that of the signal. Synchrony of the noise-
sustained oscillations in the coupled chain due to external forcing is very
similar to that of self-sustained oscillators in Chap. 3. Such array-enhanced
resonances may be important in neural systems, since coupling and noise
together can establish a much higher sensitivity to both the frequency and
the amplitude of weak external signals.
by noise. The waves die out and the media relax to the homogeneous steady
states when the noise is ceased in the simulations.
13.5 Conclusions
In this chapter, constructive effects of noise on synchronization and
synchronization-like phenomena of nonlinear systems have been discussed.
The main findings are:
14.1 Introduction
In the previous chapters we have considered synchronization of oscillators
which have regular arrangement (1D arrays or 2D lattices) with the coupling
extended to the nearest neighbors (local coupling), see Fig. 1.2, or global cou-
pling among all the osicllators (e.g., Sects. 5.5, 10.9, 13.3.3 and 13.3.4). Such
simple coupling topology of the oscillators is relevant to many experimental
and natural situations.
The locally or globally coupled networks can be considered as regular net-
works. However, there are many real-world systems which are neither locally
nor globally coupled, but often display a much more complicated coupling
topology. Examples include Internet and world-wide-web in communication
systems, neural system or genetic regulation in biology, epidemic spreading
and synchronization in social and ecological systems, etc. [495].
These systems can be characterized by complex networks. We distinguish
three main classes of complex networks [495]:
1. Random networks. About 40 years ago Erdös and Rényi (ER) [496] studied
random networks (they called them random graphs) where a pair of nodes
i and j are connected with a probability p (Fig. 14.1a). In such ER random
networks, the connection is fully random, which has the advantage that
the distance of the shortest path (path length) between any pair of nodes
is very small even for very large networks.
2. Small-world networks (SWNs). While many realistic network systems are
not regular as lattices with only local connections, they are also not fully
random as the ER networks. Especially, many of them display high clus-
tering, i.e., two connected nodes are also connected to a common third
node (thus forming a triangle), which form some local communities. In
their seminal work, Watts and Strogatz [497] proposed the small-world
network (SWN) model. They started with a regular ring of nodes, each
connected to its k nearest neighbors; then with a probability p, each link
318 14 Networks with Complex Topology
Fig. 14.1. The three basic classes of complex networks. (a) ER random networks,
(b) small-world networks (SWNs), (c) scale-free networks (SFNs)
P (k) ∼ k −γ , (14.1)
viewpoint of network topology. It has been shown that both the small-world
and the scale-free properties are universal in many real-world complex systems
(cf. [495, 503–506] and references therein).
A very important issue in the study of complex systems is the interplay
between structure and dynamics. The topology of the networks can have a
systematic influence on their physical and dynamical properties, such as error
and attack tolerance [507], percolation transition [508, 509], epidemic spread-
ing [510–512], cascading failures [513], etc. For a recent review on dynamical
process of complex networks, refer to [514].
A special aspect of this interplay is synchronization. Synchronization of
oscillators acting on the nodes is one of the widely studied dynamical behav-
ior on complex networks. A basic question is: for which network structure
(topology) we can obtain an optimal synchronization? It has been shown that
the SWNs provide a better synchrony of coupled excitable neurons in the
presence of external stimuli [515], compared with regular arrays with local con-
nections and ER random networks. In pulse-coupled oscillators, synchroniza-
tion becomes optimal in the small-world regime [516], and it is degraded when
the degree becomes more heterogeneous with increased randomness [517,518].
Investigation of phase oscillators [519] or circle maps [520] on SWNs has shown
that when more and more shortcuts are created at larger rewiring probability
p, the transition to the synchronization regime becomes easier [519]. These
observations have revealed that the ability of a network to synchronize is
generally enhanced in SWNs compared with regular chains. This enhanced
synchronization in SWNs has also been analyzed in the context of com-
plete synchronization of identical chaotic systems [521–525]. Physically, this
enhanced synchronizability was attributed to the decreasing of the average
network distance due to the shortcuts.
More recently, it has been shown that the ability of complete synchroniza-
tion of identical chaotic oscillators also depends critically on the heterogeneity
of the degree distribution [526]. In particular, random networks with strong
heterogeneity in the degree distribution, such as SFNs, are more difficult to
synchronize than random homogeneous networks [526], despite the fact that
heterogeneity reduces the average distance between the nodes [527, 528]. The
synchronizability in [521–526] is based on the linear stability of the complete
synchronization state using the spectral analysis of the network coupling matrix.
In this chapter, we discuss synchronization of nonlinear oscillators coupled
in complex networks. Our emphasis is to demonstrate how the network topol-
ogy and the connection weights influence the synchronization behavior of the
oscillators. In Sect. 14.2, we present the general dynamical equations and the
linear stability analysis for CS state when the oscillators are identical. Then
we study phase synchronization (PS) of nonidentical oscillators in SWNs
(Sect. 14.3) and SFNs (Sect. 14.4). Section 14.5 presents a qualitative analy-
sis of the hierarchical synchronization observed in Sect. 14.4. Section 14.6
is devoted to the effects of weighted coupling on synchronizability of the
networks.
320 14 Networks with Complex Topology
where JF(s) and JH(s) are the Jacobians on s. Equation (14.4) can be diag-
onalized into N decoupled blocks of the form
η̇j = [JF(s) − dΛj JH(s)] ηj , j = 1, · · · , N, (14.5)
where ηj is the eigenmode associated to the eigenvalue Λj of the coupling
matrix G. Here Λ1 = 0 corresponds to the eigenmode parallel to the synchro-
nization manifold, and the other N −1 eigenvalues Λj represent the eigenmodes
transverse to the synchronization manifold. We assume that the networks are
connected (without unconnected subnetworks), so that only one eigenvalue is
zero. The largest Lyapunov exponent (LE) λ(dΛj ) of (14.5) determines the
linear stability of the corresponding eigenmode ηj . The mode is damped if
λ(dΛj ) < 0. The synchronized state s is stable if λ(dΛj ) < 0 for j = 2, · · · , N ,
i.e., all the transverse modes are damped. Note that the same analysis can be
carried out for regular networks, as already discussed in Sect. 7.1. The above
stability analysis will be used to study CS of weighted networks in Sect. 14.6.
14.3 Phase Synchronization in Small-World Networks of Oscillators 321
d
N
ẋj = τj F(xj ) + Aji (xi − xj ), j = 1, . . . , N, (14.6)
kj i=1
50 50
(a) (b)
40 40
<Var(xj)>
30 p=0.01 30
Var(X)
p=0.1
p=0.3
20 20
p=0.5
10 10
0 0
0.0 1.0 2.0 3.0 4.0 5.0 0.0 1.0 2.0 3.0 4.0 5.0
d d
1.5
(a)
Ωj
1.0
0.5
1.5
(b)
Ωj
1.0
0.5
1.5
(c)
Ωj
1.0
0.5
0 512 1024
Node index j
Fig. 14.3. Oscillation frequency Ωj of the oscillators in SWNs (14.6) with the
shortcut probability p = 0.5 for different coupling strength. (a) d = 0, (b) d = 0.1,
and (c) d = 1.5
15 15
(a) (b)
10 10
max(x1)
max(X)
OD OD
0 0
0.0 2.0 4.0 6.0 8.0 0.0 2.0 4.0 6.0 8.0
d d
Fig. 14.4. Bifurcation diagram of the mean field (a) and of an individual oscillators
(b) against the coupling strength d in SWNs (14.6) with the shortcut probability
p = 0.5
50
40 Var(X)
<Var(xj)>
30
20
10
0
−3.0 −2.5 −2.0 −1.5 −1.0 −0.5 0.0
log10p
Fig. 14.5. Synchronization behavior vs. the shortcut probability p in SWNs (14.6)
for a fixed coupling strength (here d = 1.5)
100
2
10
10−1
P(k)
10−2
kj 10−3
101 102
101 k
Fig. 14.6. Degree sequence kj of a random SFNs (14.7) with N = 1, 000 nodes,
kmin = 5 and the scaling exponent γ = 3. The inset shows the power-law distribution
P (k) ∼ k−γ , where the flat tail results from finite size effects
2.0
1.5
Var(X)
1.0
0.5
0.0
0.0 0.2 0.4 0.6 0.8 1.0
d
Fig. 14.7. The variance of the mean field as a function of the coupling strength d
in SFNs with various scaling exponents. The symbols are: circles (γ = 3); squares
(γ = 4), and triangles (γ = ∞). The solid line is the analytically obtained results
(14.11) for globally coupled oscillators. The networks have the same mean degree
K = 10 and size N = 1, 000
d
N
ẋj = τj F(xj ) + Aji (xi − xj ), j = 1, . . . , N, (14.7)
K i=1
where we have chosen the van der Pol oscillators for F here: (ẋ = y, ẏ =
0.3(1 − x2 )y − x) and the coupling is added to both variables x and y (cf.
Sect. 2.2 for the properties of the van der Pol oscillator). Again the timescale
parameters τj follow a uniform distribution in the interval [1 − Δτ, 1 + Δτ ].
In our simulations, we set Δτ = 0.2.
Let us first examine the collective oscillations in the network. Figure 14.7
shows the variance of the mean field X as a function of the coupling strength
d for SFNs with the same mean degree K = 10, but various scaling exponents
γ = 3, 4, and ∞. It is seen that all the networks generate a coherent collective
oscillation when the coupling strength is larger than a critical value dcr ≈ 0.2.
However, networks with more heterogeneous degrees, i.e., smaller γ, generate a
326 14 Networks with Complex Topology
1.2 1.2
Ωj 1.0 1.0
(a) (b)
0.8 0.8
0 500 1000 0 500 1000
Node index j Node index j
Fig. 14.8. The oscillation frequencies Ωj of the oscillators in a SFN (14.7) with
a weak coupling strength d = 0.25. (a) Cluster synchronization regime in a hetero-
geneous network (γ = 3) and (b) global synchronization regime in a homogeneous
one (γ = ∞). The networks have the mean degree K = 10 and size N = 1, 000
0.8
1.0
0.6
0.8
ΔXj
sj
0.4
0.6 0.2
(a) (c)
0.4 0.0
0 500 1000 0 500 1000
Node index j Node index j
100
1.0 (b) (d)
0.8
ΔX(k)
s(k)
0.6 10−1
0.4 1 1 2
10 10 10 101 10
1
10
2
k k
Fig. 14.9. (a) Phase synchronization order parameter sj of node j (14.12) with
respect to the mean field X. Solid line: heterogeneous network (γ = 3); Dotted line:
homogeneous network (γ = ∞). (b) Average value s(k) of nodes with degree k as
a function of k in the heterogeneous network. (c),(d) as (a) and (b), but for the
distance ΔXj and its average value ΔX(k), respectively. The solid line in (d) with
slope α = 0.82 is plotted for reference. The results are averaged over 50 realizations of
random distribution of the time scale parameter τj . The coupling strength d = 0.25,
and N = 1, 000, and K = 10
We have also measured the distance between the state xj and the mean
field X, ΔXj = |X−xj |. In the homogeneous network, on average this measure
is also uniform for all the nodes, whereas in the heterogeneous network, it
again depends on the degree kj (Fig. 14.9c): the larger kj , the smaller ΔXj on
average. The average value ΔX(k) of nodes with degree k can be calculated
similar to (14.13) and shows a power-law dependence on k, ΔX(k) ∼ k −α ,
with α ≈ 0.82 (Fig. 14.9d). We find that the exponent α is largely independent
of γ in the degree distribution P (k) ∼ k −γ .
For the strong coupling regime (d = 1.0), the frequencies of all the oscillators
are locked mutually as well as locked to the mean field; as a result, the PS order
parameter is sj = 1 for all oscillators in both heterogeneous and homogeneous
networks, i.e., the network is globally phase synchronized. Moreover, in the
homogeneous networks, the oscillators are almost completely synchronized
in the sense that ΔXj is small and uniform on average (Fig. 14.10c). This
difference is due to a small phase difference between the oscillator and the
mean field X, Δφj = |φj − φX | (averaged over time) (Fig. 14.10a). In the
heterogeneous network, the synchronization difference ΔXj is still heterogeneous
(Fig. 14.10c), showing a dependence ΔX(k) ∼ k −α with α ≈ 0.87 (Fig. 14.10d),
14.4 Synchronization in Scale-Free Networks of Oscillators 329
0.3 0.4
(a) (c)
0.2
Δφj
ΔXj
0.2
0.1
0.0 0.0
0 500 1000 0 500 1000
Node index j Node index j
(b) (d)
10−1 10−1
ΔX(k)
Δφ(k)
10−2 10−2
1 10 100 1 10 100
k k
Fig. 14.10. (a) Average phase difference Δφj between a node j and the mean field
X. Solid line: heterogeneous network (γ = 3); Dotted line: homogeneous network
(γ = ∞). (b) Average value Δφ(k) of nodes with degree k as a function of k in the
heterogeneous network. (c),(d) as (a) and (b), but for the absolute difference ΔXj
and its average value ΔX(k), respectively. The solid lines in (b) and (d) with slope
α = 0.87 are plotted for reference. The results are averaged over 50 realizations of
the initial frequency distribution. The coupling strength d = 1.0, N = 1, 000 and
K = 10
similar to the case of weak coupling in Fig. 14.9d. Again, the synchronization
difference is induced by a heterogeneous phase difference Δφj (Fig. 14.10a).
The average phase difference displays the same power-law dependence on the
degree, Δφ(k) ∼ k −α , with the same α ≈ 0.87 (Fig. 14.10b).
In networks with completely synchronized identical oscillators, desynchro-
nization can be induced due to noise, and the degree of desynchronization
also depends on the connection degree. This is illustrated for oscillators in a
network with N = 100, γ = 3, and kmin = 5. Since the van der Pol oscillators
are nonchaotic, they become completely synchronized for a nonvanishing cou-
pling strength d > 0. Now an independent Gaussian noise from the normal
distribution N (0, 1), with amplitude D = 0.5, is added to each variable of each
oscillator in the network. This noise induces a desynchronization among the
oscillators. The oscillators still have the same average oscillation frequency,
but phase slips can occur. The de-synchronization property is not uniform in
the network, but depends on the degree k of nodes. For the coupling strength
d = 0.25, the time series x1 (t) of the oscillator with a large degree k = 44
is much cleaner compared with xN (t) of the oscillator with a small degree
k = 5, since the former is coupled stronger to the mean field X(t) accord-
ing to (14.9) and becomes more resistant to noise perturbations (Fig. 14.11).
330 14 Networks with Complex Topology
(a) 4.0
2.0
X(t) 0.0
−2.0
−4.0
(b) 4.0
2.0
x1(t)
0.0
−2.0
−4.0
4.0
(c)
2.0
xN(t)
0.0
−2.0
−4.0
0.0 100.0 200.0 300.0 400.0
time
Fig. 14.11. Time series of the mean field X(t) (a), of the node with the maximal
degree k = 44, x1 (t) (b), and of a node with the minimal degree k = 5, xN (t) (c).
The oscillators are identical with τj = 1 in (14.7). The coupling strength d = 0.25,
and N = 100 and kmin = 5
ST
j
0.8 40
20
0.7
4 10 20 40 20 40 60 80 100
k i
Fig. 14.12. (a) Average phase synchronization parameter s(k) vs. k. (b) An effec-
tive cluster; a point is plotted when sij ≥ ST . The network is the same as in
Fig. 14.11
Figure 14.12a shows s(k) as a function of k for this network. These results
indicate that nodes with large degrees form an effective cluster when they all
synchronize closer to the common mean field. Such an effective cluster is shown
in Fig. 14.12b where the bright point corresponds to sji ≥ ST = 0.82. Here
sji is the phase synchronization order parameter between a pair of oscillators
14.5 Mean-Field Analysis of Hierarchical Synchronization 331
which have the same form as (14.5), except that Λj is replaced by kj /K and
H(s) by X. Here I is the identity matrix. The largest Lyapunov exponent
λ(kj ) of (14.14) is a function of kj ,
ΔX(k) ∼ k −1 , (14.17)
which explains qualitatively the numerically observed scaling in Figs. 14.9d and
14.10d. The deviation of the scaling exponents α ≈ 0.85 from the linear result
α = 1 may result from the mean field approximation and significant nonlinearity,
so that the linear analysis in (14.16) is only a first-order approximation.
The hierarchical synchronization and effective clusters are also oberved
in network of chaotic oscillators where the coupling strength is below the
threshold of CS [538].
N
Ij = Aji Wji . (14.18)
i=1
where the weights are defined for the connections of a given network topology
and θ is a tunable parameter. Model I reproduces property (1) and has the
same weighted structure as many real networks [539]. In degree-uncorrelated
networks, Model I also reproduces the expected scaling I(k) ∼ k 1+θ . Model
II incorporates other realistic features and always reproduces property (2).
These models also include many previously studied systems as special cases.
For θ = 0, both models correspond to the widely studied case of unweighed
networks [523, 526, 559], where the weights are uniform (Wji = 1). For Model
II with θ = −1, the intensities are fully uniform (Ij = 1), as in a number of
previous studies about synchronization of coupled maps [522, 533–535].
Note that the weight matrix W is in general asymmetric for Model II.
The corresponding coupling matrix G in (14.2) is also asymmetric and can be
written as G = Dθ L, where D = diag{k1 , k2 , . . . kN } is the diagonal matrix
of the degrees and L = DI − A is the (symmetric) Laplacian matrix. Since
is valid for any Λ, we have that the spectrum of eigenvalues of the matrix G
is equal to the spectrum of a symmetric matrix defined as H = Dθ/2 LDθ/2 .
As a result, all the eigenvalues of G are real and can be ordered as [560, 561]
0 = Λ1 ≤ Λ2 · · · ≤ ΛN . (14.22)
0.3
ε1 dΛ2 dΛj dΛN ε2
0.0
Λ
−0.3
−0.6
0.0 1.0 2.0 3.0 4.0
ε
Fig. 14.13. A schematic illustration of master stability. The jth eigenmode in
(14.5) is stable when 1 < dΛj < 2 where the largest Lyapunov exponent Λ of
(14.5) is negative
1 < dΛj < 2 , where the thresholds 1 and 2 are determined only by F,
H, and s. The network is thus synchronizable for some d values when all the
transverse eigenmodes are damped (Fig. 14.13), with the following condition:
Πj ∼ kj τj−α , (14.25)
Wij=(kikj)θ Wij=kiθ
103
α = −3 102
2
R 10
101
101 (a) α=0 (b)
0 0
10 10
−3.0 −2.0 −1.0 0.0 1.0 −3.0 −2.0 −1.0 0.0 1.0
θ θ
Fig. 14.14. Eigenratio R (14.24) as functions of θ for Model I (a) and Model II (b)
in growing SFNs with the aging exponent α = 0 (◦) and α = −3 (•). Each symbol
is an average over 50 realizations of the networks, for K = 20 and N = 210 . The
solid lines are the approximations of R by (14.36) with AR = 0.47
102
Wij = 1
R Wij = (kikj)1
101
Wij = ki1
100
−3.0 −2.0 −1.0 0.0
α
Fig. 14.15. Eigenratio R (14.24) vs.α for growing SFNs with θ = 0 (circles) and
θ = −1 in Model I (triangles) and in Model II (dots). Dotted line: RH (K) in (14.35).
Solid lines: approximation by (14.36) with AR = 0.47. The other parameters are
the same as in Fig. 14.14
N
H̄W
j = (1/Ij ) Wji Aji H(xi ) = (kj /Ij ) Wji H(xi ) j (14.28)
i=1
as the weighted local mean field of all the neighbors connected to the oscil-
lator j, where j denotes average over the kj neighbors of the node j.
Equation (14.2) then can be rewritten as
j − H(xj )].
ẋj = F(xj ) + dIj [H̄W (14.29)
where
N
H̄j = (1/kj ) Aji H(xi ) (14.31)
i=1
is the unweighted local mean field. Note that in Model II, H̄W j ≡ H̄j since
Wji = Ij /kj . If the network is sufficiently random with a large enough mini-
mum degree kmin , the local mean field H̄j in (14.31) can be approximated by
the global mean field of the network, H̄j ≈ H̄. Moreover, for small perturba-
tions close to the synchronized state s, we may assume H̄j ≈ H(s), and the
system is approximated as
indicating that the oscillators are decoupled and forced by a common oscillator
ṡ = F(s), with the forcing strength being proportional to the intensity Ij . If
there exists some d satisfying ε1 < dIj < ε2 for all j, then all the oscillators
are synchronizable by the common driving H(s), corresponding to CS of the
whole network.
This finding suggests that the eigenratio R can be approximated as
where Imax and Imin are the maximal and minimal values of the intensities,
respectively. For fully uniform intensities in Model II at θ = −1, R = 1 by this
14.6 Synchronization Properties of Weighted Networks 337
As shown in Fig. 14.15 by the dotted line, (14.35) provides a good approxi-
mation under the weaker condition kmin 1, when the intensity is uniform
Ij = 1 (Model II, θ = −1), regardless of the degree distributions at different
α values. Physically, the dependence on the mean degree K can be under-
stood as due to deviations or fluctuations of the local mean field H̄j from
the global mean field H̄. This effect is similar when the intensities are not
uniform. Thus, we may assume that the contribution due to the number of
connections is statistically independent of the contribution due to the strength
of the connections. Accordingly, for general weighted random networks with
arbitrary distributions of Ij and kj , we get from (14.33) and (14.35)
Imax
R = AR RH (K), (14.36)
Imin
where AR is a constant of the order of 1.
With a single parameter AR = 0.47, (14.36) approximates the eigenratio R
very closely for different networks and weighted coupling schemes, including
the unweighed networks (θ = 0) (Figs. 14.14 and 14.15). The fitting parameter
underestimates R slightly only when the intensities become rather homoge-
neous (Imax /Imin < 3).
These results demonstrate that the synchronizability in random complex
networks is determined by two leading parameters, the mean degree K and
the heterogeneity of the intensity Ij as measured by the ratio Imax /Imin . This
dependence is universal for networks with different degree distribution. More
analysis also shows that it is also universal for networks displaying nontrivial
clustering and degree correlations [565].
The above analysis on identical oscillators can serve as a good approxi-
mation even when the oscillators are not fully identical. Now we consider PS
in weighted complex networks of nonidentical van der Pol oscillators as in
Sect. 14.4:
338 14 Networks with Complex Topology
2.0
θ=0
1.0 θ = −1
0.0
X
−1.0
(a) (b)
−2.0
0 20 40 60 80 100 0 20 40 60 80 100
time time
Fig. 14.16. Time series of the mean field X in random SFNs of nonidentical van der
Pol oscillators with γ = 3. The coupling strengths are (a) d = 0.1 and (b) d = 0.2
d
N
ẋj = τj F(xj ) + Aji Wji (xi − xj ), j = 1, . . . , N, (14.37)
I ∗ i=1
N
where I ∗ = (1/N ) j Ij is the mean value of the intensity. Note that for
the unweighted case, Wji = 1, and I ∗ is just the mean degree, i.e., I ∗ = K,
so that (14.37) becomes the same as (14.7). Here again we assume that the
parameters τj follow a uniform distribution in the interval [1 − Δτ, 1 + Δτ ].
In our simulations we set Δτ = 0.2. The SFNs are generated as in Sect. 14.4
with γ = 3, kmin = 10 (K = 20), and N = 1, 024. In Fig. 14.16, we show time
series of the mean field X for unweighted (θ = 0) and weighted random SFNs
(Model II, θ = −1). For small coupling strength d, neither of the networks
display significant collective behavior (Fig. 14.16a). As the coupling strength
is increased, coherent collective oscillations emerge for both weighted and
unweighted networks, but the oscillations are much more pronounced for the
networks with θ = −1 (Fig. 14.16b).
In Fig. 14.17, we show the variance of the mean field X as a function of d.
The variance is approximately zero for small coupling strength, increases
sharply as d is increased beyond a certain critical value, and saturates
for large d (Fig. 14.17). The overall behavior is similar for weighted and
unweighted networks, but, again, the variance is significantly larger for net-
works with θ = −1 at the same value of d. Moreover, the variance of random
SFNs for θ = −1 is well approximated by the variance of random homoge-
neous networks with the same mean degree (Fig. 14.17), which are networks
that exhibit good phase synchronization properties as shown in Sect. 14.4. In
both the cases, the intensities Ij are fully uniform. These results confirm that
PS in random networks is also universally determined by the mean degree
K and the distribution of the intensity Ij . Synchronization is enhanced in
networks with more homogeneous intensities.
14.7 Conclusions 339
2.0
1.5
Var(X) 1.0
0.5
0.0
0.0 0.2 0.4 0.6 0.8 1.0
d
Fig. 14.17. Variance of the mean field X as a function of d for random SFNs
of van der Pol oscillators with θ = 0 (circles) and θ = −1 (dots). The diamonds
correspond to random homogeneous networks with the same mean degree (K = 20)
of the random SFNs. The results are averaged over 20 realizations
14.7 Conclusions
In this chapter synchronization of nonlinear oscillators on complex networks
has been discussed. We have shown that the topology as well as the weight
of the connections has significant influences on the synchronization of the
networks. The main results are:
– In SWNs, the synchronizability of the networks is enhanced significantly
by adding some shortcuts to the underlying locally coupled networks.
Oscillation death and global synchronization of SWNs have been demon-
strated.
– In SFNs displaying heterogeneous degrees, the synchronization behavior
of a node depends on its degree in unweighted networks where the coupling
strength is uniform for all the connections in the network. Hubs (nodes
with large degrees) mainly contribute to the mean field of the network
and are dominant in forming an effective synchronization cluster.
– Weighted coupling influences significantly the synchronizability of com-
plex networks. In random networks, the synchronizability is universally
controlled by the mean degree and the heterogeneity of the intensities
of nodes. Networks with more homogeneous intensities display enhanced
synchronization behavior.
Synchronization of complex networks is relevant in many real-world sys-
tems. For example, brain networks [543, 544] display a hierarchy of oscillation
and synchronization on various spatial and temporal scales [545], and syn-
chronization of distributed brain activity has been proposed as an important
mechanism for neural information processing [546–548]. The analysis of the
anatomical connectivity of the mammalian cortex [549] and the functional con-
nectivity of the human brain [550–552] have shown that both share typical
340 14 Networks with Complex Topology
20. F.C. Hoppensteadt and E.M. Izhikevich, Weakly Connected Neural Networks,
Springer-Verlag, New York, 1997.
21. C.V. Gray, J.Computat.Neuroscience 1, 11 (1995).
22. W. Singer and C.M. Gray, Ann. Rev. Neorosci. 18, 555 (1995).
23. C.J. DeLuca, A.M. Roy, and Z. Erim, J. Neurophysiol. 70, 2010 (1993).
24. W. Singer, Ann. Rev. Physiol. 55, 349 (1993).
25. P.A. Tass Phase resetting in Medicine and Biology Springer, Berlin, 1999.
26. R.K. Wang, R.D. Traub, and R. Miles, Adv. Neurol. 44, 583 (1986).
27. J.M. Gonzalez-Miranda Synchronization and Control of Chaos. An Introduc-
tion for Scientists and Engineers World Scientific, Singapore, 2004.
28. C.W. Wu Synchronization in coupled chaotic circuits and systems World
Scientific, Singapore, 2002.
29. M.E. Yalcin, J.A.K. Suykens, and J.P.L. Vandewalle, Cellular Neural Net-
works, Multi-Scroll Chaos and Synchronization World Scientific, Singapore,
2005.
30. M.K. McClintock, Nature 229, 244 (1971).
31. B. Blasius, A. Huppert, and L. Stone, Nature 399, 354–359 (1999).
32. D.E. Postnov, A.G. Balanov, and E. Mosekilde, Advances in Complex Systems
1, 181 (1998).
33. R.A. York and R.C. Compton IEEE Trans. Microwave Theory Tech. 39, 1000
(1991).
34. K. Wiesenfeldt, P. Colet, and S.H. Strogatz, Phys. Rev. Lett. 76, 404 (1996).
35. K. Wiesenfeldt, P. Colet, and S.H. Strogatz, Phys. Rev. E 57, 1563 (1998).
36. A. Rodriguez-Angeles and H. Nijmeijer, Int. J. Control 74, 1311 (2001).
37. Y.-H. Liu, Y. Xu, and M. Bergerman, IEEE Trans. Robot. Autom. 15, 258
(1999).
38. Z. Jiang and M. McCall, J. Opt. Soc. Am. 10, 155 (1993).
39. S.Yu. Kourchatov, V.V. Likhanskii, A.P. Napartovich, F.T. Arecchi, and
A. Lapucci, Phys. Rev. A 52, 4089 (1995).
40. A.F. Glova, S.Yu. Kurchatov, V.V. Likhnaskii, A.Yu. Lysikov, and
A.P. Napartovich, Quant. Electron. 26(6), 500 (1996).
41. Y. Tanguy, J. Houlihan, G. Huyet, E.A. Viktorov, and P. Mandel, Phys. Rev.
Lett. 96, 053902 (2006).
42. S. Peles, J.L. Rogers, and K. Wiesenfeld, Phys. Rev. E 73, 026212 (2006).
43. D. Tsygankov and K. Wiesenfeld, Phys. Rev. E 73, 026222 (2006).
44. S. Yanchuk, A. Stefanski, T. Kapitaniak, and J. Wojewoda, Phys. Rev. E 73,
016209 (2006).
45. L. Pecora, Editor. A focus issue on synchronization in chaotic systems. Chaos
7 (1997).
46. J. Kurths, Editor. A focus issue on phase synchronization in chaotic systems.
Int. J. Bifurc. and Chaos, 10, (2000).
47. A.A. Andronov, A.A. Vitt, and S.E. Khaykin, Theory of Oscillations,
Gostekhizdat, Moscow, 1937. (In Russian); English Translation: Pergamon
Press, Oxford - NY - Toronto, 1966.
48. B. van der Pol, Phil. Mag. 3, 64 (1927).
49. N. Minorsky Nonlinear Oscillations Van Nostrand, Princeton, NJ, 1962.
50. M.I. Rabinovich and D.I. Trubetskov, Oscillations and Waves in Linear and
Nonlinear Systems, Kluwer Academic Publishers, Dordrecht-Boston-London,
1989.
References 347
83. M. Steirade, D.A. McCormick, and T.J. Sejnowski, Science 262, 679 (1993).
84. F. Amzica and M. Steirade, Electroencephalogr. Clin. Neurophysiol. 2, 69
(1998).
85. W. Schultz, J. Neurophysiol. 80, 1 (1998).
86. H.D. Abarbanel et al., Phys. Usp. 39, 337 (1996) [Usp. Fiz. Nauk 166, 363
(1996).].
87. E.M. Izhikevich, IEEE Trans. on Neural Networks, (2004).
88. H.D. Abarbanel et al., Neural Comput. 8, 1567 (1996).
89. R.D. Pinto et al., Phys. Rev. E 62, 2644 (2000).
90. R.C. Elson et al., Phys. Rev. Lett. 81, 5692 (1998).
91. N.F. Rulkov, Phys. Rev. Lett. 86, 183 (2001).
92. R. FitzHugh, Biophys. J. 1, 445 (1961).
93. J. Nagumo, S. Arimoto, and S. Yoshizawa, Proc. IRL 50, 2061 (1960).
94. C.H. Luo and Y. Rudy, Circ. Res. 68, 1501 (1991).
95. I.S. Aranson, A.V. Gaponov-Grekhov, M.I. Rabinovich, A.V. Rogalsky, and
R.Z. Sagdeev, “Lattice models of nonlinear dynamics of nonequilibrium
media” (Preprint: Inst. of Appl. Phys., Gorky.) 163 (1987).
96. S.H. Strogatz Sync. The Emerging Science of Spontaneous Order. Theia
(Hyperion), New York, 2003.
97. L. Glass and M.C. Mackey, From Clocks to Chaos: The Rhythms of Life.
Princeton University Press, Princeton, NJ, 1988.
98. R.L. Stratonovich, Topics in the Theory of Eandom Noise. Gorddon and
Breach, New York, 1963.
99. A.N. Malakhov, Fluctuations in Self-Oscillatory Systems., Nauka, Moscow,
1968 (in Russian).
100. S. Boccaletti, J. Kurths, G. Osipov, D.L. Valladares, and C.S. Zhou, Physics
Reports 366, 1 (2002).
101. V.S. Anishchenko and T.E. Vadivasova, Synchronziation of Self-Oscillations
and Noise-Induced Oscillations. J. of Communications Technology and Elec-
tronics 47, 117 (2002).
102. E.V. Appleton, Proc. Cambridge Phil. Soc. (Math and Phys. Sci.) 21, 231
(1922).
103. A.A. Andronov and A.A. Vitt, Zhurnal prikladnoj Fiziki (J. Apll. Phys.) 7, 3
(1930) (in Russian).
104. A.A. Andronov and A.A. Witt, Archiv für Elektrotechnik 24, 99 (1930).
105. L.I. Mandelshtam and N.D. Papaleksi, Collected works by L. I. Mandelshtam,
2 (Izd. Akademii Nauk, Moscow), 13 (1947) (in Russian).
106. M.L. Cartwright and J.E. Littlewood, J. London Math. Soc., 20, 180 (1945).
107. M.L.Cartwright, J. Inst. Elec. Eng., 95, 88 (1948).
108. N.N. Bogolyubov and Yu.A. Mitropolsky, Asymptotic Methods in the Theory
of Nonlinear Oscillations, Gordon and Breach, New York, 1961.
109. E.F. Stone, Phys. Lett. A 163, 367 (1992).
110. A.S. Pikovsky, M.G. Rosenblum, G.V. Osipov, and J. Kurths, Physica D, 104,
219 (1997).
111. K. Josić and D.J. Mar, Phys. Rev. E 64, 066234 (2001).
112. M.A. Zaks, E.-H. Park, M.G. Rosenblum, and J. Kurths, Phys. Rev. Lett. 82,
4228 (1999).
113. E.H. Park, M. Zaks, and J. Kurths, Phys. Rev. E 60, 6627 (1999).
114. M.A. Zaks, E.-H. Park, and J. Kurths, Int. J. Bifurcation and Chaos, 10, 2649
(2000).
References 349
115. C. Schafer, M.G. Rosenblum, J. Kurths, and H.H. Abel, Nature 392, 239
(1998).
116. C. Schäfer, M.G. Rosenblum, H.H. Abel, and J. Kurths, Phys. Rev. E 60, 857
(1999).
117. A. Pikovsky, G. Osipov, M. Rosenblum, M. Zaks, and J. Kurths, Phys. Rev.
Lett., 79, 47, (1997).
118. E.R. Rosa, E. Ott, and M.H. Hess, Phys. Rev. Lett. 80, 1642 (1998).
119. A.S. Pikovsky, M. Zaks, M. Rosenblum, G. Osipov, and J. Kurths, Chaos, 7,
680 (1997).
120. A.S. Pikovsky and P. Grassberger, J. Phys. A 24, 4587 (1991).
121. D.Y. Tang, M.Y. Li, and C.O. Weiss, Phys. Rev. A 46, 676 (1992).
122. M. Arjona, J. Pujol, and R. Corbalan, Phys. Rev. A 50, 871 (1994).
123. A.M. Kulminskii and R. Vilasecar, J. Mod. Optic 42, 2295 (1995).
124. N.W. Mureithi, M.P. Paidoussis, and S.J. Price, J. Fluid Struct. 8, 853(1994).
125. H. Okamoto, N. Tanaka, and M. Naito, J. Phys. Chem. 102, 7353 (1998).
126. R. Richter, J. Peinke, and A. Kittel, Europhys. Lett. 36, 675 (1996).
127. D. Gillet, Astron. Astrophys. 259, 215 (1992).
128. D.L. Feng et al., Phys. Rev. E 54, 2839 (1996).
129. J.H. Cho et al., Phys Rev. E 65, 036222 (2002).
130. J.K. Bhattacharjee and K. Banerjee, Phys. Rev. A, 29 2301 (1984).
131. S.Y. Kim, Phys. Rev. E, 59, 2887 (1999).
132. L.D. Landau and E.M. Lifshitz Mechanics, Pergamon Press Ltd., 1976.
133. Similar equation appears in studies of the thermal convection between hori-
zontal plates when periodic driving force is applied. Behavior of the system
near the convictive threshold was studied in G. Ahlers, P.C. Hohenberg and
M. Lücke, Phys. Rev. A, 32 3493 (1985).
134. A.R. Yehia, D. Jeandupeux, F. Alonso, and M.R. Guevara, Chaos, 9, 916,
1999.
135. V.S. Afraimovich and L.P. Shilnikov In Methods of Qualitatively Theory of
Differential Equations, Gorky, 3, 1983. (In Russian); English translation Amer.
Moth. Soc. Transl. (2) 149, 201 (1991).
136. G.I. Dykman, P.S. Lnada, and Yu.I. Neimark Chaos, Solitons and Fractals,
1(4) 339 (1991).
137. C. Hugenii, Horoloquim Oscilatorium, Parisiis, France, 1673.
138. A. Mayer, Proc. Gorky State University, 2, 3 (1935) (in Russian).
139. V.I. Gaponov, Zh. Tech. Fiz. (J. Tech. Phys.) 6, 5 (1936) (in Russian).
140. A.S. Bremsen and I.S. Feinberg, Zh. Tech. Fiz. (J. Tech. Phys.) 11 959 (1941)
(in Russian).
141. K.F. Teodorchik, Dokl. Akad. Nauk SSSR (Sov. Phys. Dokl.), 40, 63 (1943)
(in Russian).
142. D.W. Storti and R.H. Rand, Int. J. Non-Linear Mech., 17, 3, 143 (1982).
143. R.H. Rand and P.J. Holmes, Int. J. Non-Linear Mech., 15, 387 (1980).
144. G.B. Ermentrout, Physica D, 41, 219 (1990).
145. D.G Aroson, G.B. Ermentrout, and N. Kopell, Physics D, 41, 403 (1990.)
146. M. Poliashenko, S.R. McKay, and C.W. Smith, Phys. Rev. A, 44, 3452 (1991).
147. H.G. Schuster and P. Wagner, Prog. Theor. Phys., 81, 939 (1989).
148. D.V.R. Reddy, A. Sen, and G.L. Jhonston, Phys. Rev. Lett., 80, 5109 (1998).
149. S.H. Strogatz, Nature (London), 394, 316 (1998).
150. A. Takamatsu, T. Fujii, and I. Endo, Phys. Rev. Lett., 85, 2026 (2000).
350 References
151. M. Dhamala, V.K. Jirsa, and M. Ding, Phys. Rev. Lett., 92, 074104 (2004).
152. R.H. Rand, A.H. Cohenm, and P.J. Holmes, in Neural Control and Rhythmic
Movements in Vertebrates, A.H. Cohen, S. Grillner, and S. Rossignol, eds.,
MIT Press, Cambridge, MA, 1989.
153. E.M. Izhikevich, SIAM J. Appl. Math. 60, 1789 (2000).
154. M.V. Ivanchenko, G.V. Osipov, V.D. Shalfeev, and J. Kurths, Physica D, 189,
8 (2004).
155. T. Chacraborty and R.H. Rand, Int. J. Non-Linear Mech., 23, (1988), 369–376.
156. H. Fujisaka and T. Yamada, Prog. Theor. Phys. 69, 32 (1983).
157. V.S. Afraimovich, N.N. Verichev, and M.I. Rabinovich, Izvestia-Vysshikh-
Uchebnykh-Zavedenii-Radiofizika 29 (9), 1050 (1986).
158. L.M. Pecora and T.L. Carroll, Phys. Rev. Lett. 64, 821 (1990).
159. K. Cuomo and A.V. Oppenheim, Phys. Rev. Lett. 71, 65 (1993).
160. J.Y. Chen, K.W. Wong, L.M. Cheng, and J.W. Shuai, Chaos, 13, 508 (2003).
161. A.S. Dmitriev and A.I. Panas, Dynamical chaos: new information carrier in
communication systems, Fizmatlit, Moscow, 2002 (In Russian).
162. A.S. Pikovsky, Sov. J. Commun. Technol. Electron. 30, 85 (1985).
163. M.G. Rosenblum, A.S. Pikovsky, and J. Kurths, Phys. Rev. Lett. 76, 1804
(1996).
164. U. Parlitz et al., Phys. Rev. E 54, 2115 (1996).
165. G.V. Osipov, A.S. Pikovsky, M.G. Rosenblum, and J. Kurths, Phys. Rev. E
55, 2353 (1997).
166. M.G. Rosenblum, A.S. Pikovsky, and J. Kurths, Phys. Rev. Lett. 78, 4193
(1997).
167. N.F. Rulkov, M.M. Sushchik, L.S. Tsimring, and H.D.I. Abarbanel, Phys. Rev.
E 51, 980 (1995).
168. L. Kocarev and U. Parlitz, Phys. Rev. Lett. 76, 1816 (1996).
169. J.Y. Chen et al., Phys. Rev. E 64, 016212 (2001).
170. K.J. Lee, Y. Kwak, and T.K. Limm, Phys. Rev. Lett. 81, 321 (1998).
171. W.-H. Kye and C.-M. Kim, Phys. Rev. E 62, 6304 (2000).
172. B. Hu, G.V. Osipov, H.-L. Yang, and J. Kurths, Phys. Rev. E, 67 066216
(2003).
173. N.F. Rulkov, Chaos, 6, 262 (1996).
174. P. Ashwin, J. Buescu, and I. Stewart, Phys. Lett. A, 193, 126 (1994).
175. T. Endo, J. Franklin Institute, B 331, 1859 (1994)
176. A. Sato and T. Endo, Proc. NDES’94, 117 (1994).
177. A.K. Kozlov and V.D. Shalfeev, Proc. NDES’95, 233 (1995).
178. N. Smith, C. Crowley, and M. Kennedy, Proc. NDES’96.
179. M.V. Korzinova, V.V. Matrosov, and V.D. Shalfeev, Int. J. Bifurcation and
Chaos, 9, 963 (1999).
180. V.D. Shalfeev and G.V. Osipov, Proc. NDES’97, 139, (1997).
181. H.L. Yang, Phys. Rev. E 63, 026213 (2001).
182. V.V. Matrosov and V.D. Shalfeev, Radiophys. Quantum Electron. 41, 1033
(1998).
183. G.V. Osipov and J. Kurths, Phys. Rev. E65, 016216 (2002).
184. G.V. Osipov, A.S. Pikovsky, and J. Kurths, Phys. Rev. Lett. (2002).
185. F.C. Hoppensteadt and E.M. Izhikevich, Weakly Connected Neural Networks,
Springer-Verlag, New York, 1997.
186. N. Kopell and G.B. Ermentrout, Common. Pure Appl. Math. 39, 623 (1986).
References 351
187. G.B. Ermentrout and N. Kopell, J. Math. Biol. 29, 195 (1991).
188. H. Takana, A. Lichtenberg, and S. Oishi, Phys. Rev. Lett 78, 2104 (1997).
189. K. Tsang and K. Nagi, Phys. Rev. E 54, R3067 (1996).
190. J. Rogers and L. Wille, Phys. Rev. E 54, R2193 (1996).
191. S. Strogatz, R. Mirollo, and P. Matthews, Phys. Rev. Lett 68, 2730 (1992).
192. H. Daido, Phys. Rev. Lett 77, 1406 (1996).
193. H. Daido, Phys. Rev. Lett 78, 1683 (1997).
194. S. Watanabe, H. Zant, S. Strogatz, and T. Orlando, Physica D 97, 429 (1996).
195. H. van der Zant, T. Orlando, S. Watanabe, and S. Strogatz, Phys. Rev. Lett
74, 174 (1995).
196. A. Ustinov, M. Cirillo, and B. Malomed, Phys. Rev. B 47, 8357 (1993).
197. A. Ustinov, M. Cirillo, B. Larsen, V.A. Gbozunov, P. Carelli, and G. Rotoli
B. Malomed, Phys. Rev. B 51, 3081 (1995).
198. Z. Zheng, B. Hu, and G. Hu, Phys. Rev. E 57, 1139 (1998).
199. M. Antoni and S. Ruffo, Phys. Rev. E 52, 2361 (1995).
200. R. Fishman and D. Stroud, Phys. Rev. B 38, 290 (1988).
201. C. Yokoi, L. Thang, and W. Cou, Phys. Rev. B 37, 2173 (1988).
202. P. Liao and R.A. York, IEEE Trans. on Microwave Theory and Techniques
41, 1810 (1993).
203. T. Heath, K. Wiesenfeldt, and R.A. York, Int. J. Bifurc. and Chaos 10, 2619
(2000).
204. M. Gabbay, M.L. Larsen, and L.S. Tsimring, Phys. Rev. E 70, 066212 (2004).
205. D. Topaj and A. Pikovsky, Physica D, 170, 118 (2002).
206. G.B. Ermentrout and N. Kopell, I, SIAM J. Math. Anal. 15, 215 (1984).
207. Z. Zheng, B. Hu, and G. Hu, Phys. Rev. Let. 81, 5318 (1998).
208. Z. Zheng, B. Hu, and G. Hu, Phys. Rev. E 62, 402 (2000).
209. H.F. El-Nashar, Y. Zhang, H.A. Cerdeira, and F. Ibiyinka, Chaos 13, 1216
(2003).
210. A.V. Gaponov-Grekhov, M.I. Rabinovich, and I.M. Starobinetz, JETP Lett.,
39, 561 (1984) [in Russian].
211. G.V. Osipov, Radiophysics 31, 624 (1988) [in Russian].
212. I.S. Aranson, A.V. Gaponov-Grekhov, M.I. Rabinovich, and I.M. Starobinetz,
JETP 90, 1707 (1986) [in Russian].
213. V.S. Anishchenko, I.S. Aranson, D.E. Postnov, and M.I. Rabinovich, Acad. of
Sci. USSR Papers 286, 1120 (1986) [in Russian].
214. V.S. Anishchenko, D.E. Postnov, and M.A. Safonova, JETP Lett., 11, 1505
(1985) [in Russian].
215. S.P. Kuznetsov and A.S. Pikovsky, Radiophysics 28, 308 (1985) [in Russian].
216. V.S. Afraimovich, M.I. Rabinovich, and V.I. Sbitnev, JETP Lett., 11, 338
(1985) [in Russian].
217. I.S. Aranson, A.V. Gaponov-Grekhov, and M.I. Rabinovich, Physica D, 33, 1
(1988).
218. D. Ruelle and F. Takens, Comm. Math. Phys., 20, 167 (1979).
219. S. Newhouse, D. Ruelle, and F. Takens, Comm. Math. Phys., 64, 35 (1979).
220. A.J. Lichtenberg and M.A. Liberman, Regular and stochastic motion,
(Springer-Verlag, New York - Heidelberg - Berlin), 1983.
221. A.T. Winfree and J. Theoret. Biol. 16, 15 (1967).
222. Y. Kuramoto, in Proc. of Int. Symp. on Mathematical Problems in Theoretical
Physics, H. Araki, ed., Lecture Notes in Physics 39 (Springer, New York,
1975).
352 References
223. S.H. Strogatz, D.M. Abrams, A. McRobie et al., Nature 438, 7064 (2005).
224. K. Lonngren and A. Scott (eds.) Solitons in Action. (Acad. Press, New York,
San Francisco, London) (1978).
225. J.S. Eilbeck, P.S. Lomdahl, O.H. Olsen, and M.R. Samuelsen, J. of Appl.
Phys., 57, 861 (1985).
226. O.H. Olsen, P.S. Lomdahl, A.R. Bishop, and I.S. Eilbeck, J. Phys. C.: Solid
State Phys. 18, 54 (1985)
227. A.R. Bishop, M.G. Forest, D.W. McLaughlin, and E.A. Overmann, Physica
D, 23, 293 (1986).
228. A.R. Bishop and P.S. Lomdahl, Physica D, 18, 54 (1986).
229. P.M. Marcus and Y. Imry Solid. State Comm. 33, 345 (1980).
230. L.Sh. Tsimring, Phys. Rev. E 48, 3446 (1993).
231. L.N. Korzinov and M.I. Rabinovich, Izv. VUZov, Applied Nonlin. Dynam. 2,
59 (1994) [in Russian].
232. M. Bazhenov, M. Rabinovich, and L. Rubchinsky, J. Stat. Phys. 83, 1165
(1996).
233. Y. Braiman, J.F. Lindner, and W.L. Ditto, Nature 378, 465 (1995).
234. Y. Braiman, W.L. Ditto, K. Wiesenfeld, and M.L. Spano, Phys. Lett. A 206,
54 (1995).
235. N.E. Diamant, P.K. Rose, and E.J. Davidson, Amer. J. Physiol. 219, 1684
(1970).
236. A.B. Zobnin, M.I. Rabinovich, and M.M. Sushchik, Izv. Akad. Nauk SSSR,
ser. Fizika Atmosfery i Okeana 26, 1289 (1990) [in Russian].
237. S.K. Sarna, E.E. Daniel, and Y.I. Kinoma, Amer. J. Physiol. 221, 166 (1971).
238. D. Robertson-Dunn and D.A. Linkens, Med. Biol. Engng. 12, 750 (1974).
239. B.H. Brown, H.L. Duthie, A.R. Horn, and R.H. Smallwood, Amer. J. Physiol.
229, 384 (1975).
240. R.J. Patton and D.A. Linkens, Med. Biol. Engng. Computing 16, 195 (1978).
241. S.D. Drendel, N.P. Hors, and V.A. Vasiliev, in Dynamics of Cell Popula-
tions (Nizhny Novgorod University Press, Nizhny Novgorod, 1984), p. 108
[in Russian].
242. V.A. Vasiliev, Yu.M. Romanovsky, and V.G. Yakhno, Autowave Processes
(Nauka, Moscow, 1987) [in Russian].
243. Y. Yamaguchi and H. Shimizu, Physica D 11, 212 (1984).
244. K. Bar-Eli, Physica D 14, 242 (1985).
245. G.B. Ermentrout, in Nonlinear Oscillations in Biology and Chemistry. Lecture
Notes in Biomathematics, Vol. 66, ed. H. Othmer (Springer, Berlin, 1986).
246. G.B. Ermentrout and W.C. Troy, SIAM J. Appl. Math. 39, 623 (1986).
247. I.S. Aronson and L. Kramer, Rev. Mod. Phys. 74, 99 (2002).
248. D. Somers and N. Kopell, Physica D 89, 169 (1995).
249. T.E. Vadivasova, G.I. Strelkova, and V.S. Anishchenko, Phys. Rev. E 63,
036225 (2001).
250. D. Hansel and H. Sompolinsky, Phys. Rev. Lett. 68, 718 (1992).
251. D. Hansel, Int. J. Neural Sys. 7, 403 (1996).
252. F. Pasemann, Physica D 128, 236 (1999).
253. P. Varona, J.J. Tores, H.D.I. Abarbanel, M.I. Rabinovich, and R.C. Elson,
Biol. Cybern. 84, 91 (2001).
254. G. Balazsi, A. Cornell-Bell, A.B. Neiman, and F. Moss, Phys. Rev. E 64,
041912 (2001).
References 353
370. Y. Braiman, W.L. Ditto, K. Wiesenfeld, and M.L. Spano, Phys. Lett. A 206,
54 (1995).
371. N. Mousseau, Phys. Rev. Lett. 77, 968 (1996).
372. Y. Braiman, J.F. Lindner, and W.L. Ditto, Nature 378, 465 (1995).
373. J.F. Lindner, B.S. Prusha, and K.E. Klay, Phys. Lett. A 231, 164 (1997).
374. A. Gavrielides, T. Kottos, V. Kovanis, and G.P. Tsironis, Phys. Rev. E 58,
5529 (1998).
375. J.F. Hirsch, B.A. Huberman, and D.J. Scalpiano, Phys. Rev. A 25, 519 (1982).
376. S.C. Venkataramani, B.R. Hunt, and E. Ott, Phys. Rev. E 54, 1346 (1996).
377. W. Wiesenfeld and F. Moss, Nature 373, 33 (1995).
378. A.R. Bulsara and L. Gammaitoni, Phys. Today 49, 39 (1996).
379. L. Gammaitoni, P. Hanggi, P. Jung, and F. Marchesoni, Rev. Mod. Phys. 70,
223 (1998).
380. V.S. Anishchenko, A.B. Neiman, F. Moss, and L. Schimansky-Geier, Physics–
Uspekhi 42, 7 (1999).
381. J.F. Lindner, B.K. Meadows, W.L. Ditto, M.E. Inchiosa, and A.R. Bulsara,
Phys. Rev. E 53, 2081 (1996).
382. M. Locher, G.A. Johnson, and E.R. Hunt, Phys. Rev. Lett. 77, 4698 (1996).
Phys. Rev. E 58, 2843 (1998).
383. J.F. Lindner, S. Chandramouli, A.R. Bulsara, M. Locher, and W.L. Ditto,
Phys. Rev. Lett. 81, 5048 (1998).
384. M. Locher, D. Cigna, and E.R. Hunt, Phys. Rev. Lett. 80, 5212 (1998).
385. S. Kadar, J. Wang, and K. Showalter, Nature 391, 770 (1998).
386. P. Jung and G. Mayer-Kress, Phys. Rev. Lett. 74, 2130 (1995).
387. H. Hempel, L. Schimansky-Geier, and J. Garcia-Ojalvo, Phys. Rev. Lett. 82,
3713 (1999).
388. A. Zaikin, J. Garcia-Ojalvo, L. Schimansky-Geier, and J. Kurths, Phys. Rev.
Lett. 88, 010601 (2002).
389. E.I. Volkov, E. Ullner, A.A. Zaikin, and J. Kurths, Phys. Rev. E 68, 026214
(2003).
390. H. Zhonghuai, Y. Lingfa, X. Zuo, and X. Houwen, Phys. Rev. Lett. 81, 2854
(1998).
391. M.A. Santos and J.M. Sancho, Phys. Rev. E 59, 98 (1999).
392. A.T. Stamp, G.V. Osipov, and J.J. Collins, Chaos, 12, 931 (2002).
393. Chaos, Focus Issue: Mapping and Control of Complex Cardiac Arrhythmias,
12, Editors: L. Glass and D. Christini, 3 (2002).
394. American Heart Association, Heart Disease and Stroke Statistics-2003 Update
(American Heart Association, Dallas, 2002).
395. D.L. Hayes, M.A. Lloyd, and P.A. Friedman, Cardiac Pacing and Defibrilla-
tion: A Clinical Approach (Futura Pub Co, 2000.)
396. J. Jalife and R.A. Gray, Acta. Physiol. Scand., 157, 123 (1996)
397. J. Jalife, R.A. Gray, G.E. Morley, and J.M. Davidenko, Chaos, 8, 79 (1998).
398. A.V. Panfilov, Chaos, 8, 57 (1998).
399. F.X. Witkowski, L.J. Leon, P.A. Penkoske, W.R. Giles, M.L. Spano, W.L.
Ditto, and A.T. Winfree, Nature, 392, 78 (1998).
400. L. Tung, Proc. IEEE., 84, 366 (1996).
401. O. Steinbock, J. Schütze, and S.C. Müller, Phys. Rev. Lett., 68, 248 (1992).
402. V.N. Biktashev and A.V. Holden, Chaos Solitons Fractals, 5, 575 (1995).
403. R.-M. Mantel and D. Barkley, Phys. Rev. E, 54, 4791 (1996).
References 357
404. S. Grill, V.S. Zykov, and S.C. Müller, Phys. Rev. Lett., 75, 3368 (1995).
405. V.S. Zykov, A.S. Mikhailov, and S.C. Müller, Phys. Rev. Lett., 78, 3398
(1997).
406. A.M. Pertsov, J.M. Davidenko, R. Salomonsz, W.T. Baxter, and J. Jalife,
Circ. Res., 72, 631 (1993).
407. G.V. Osipov, B.V. Shulgin, and J.J. Collins, Phys. Rev. E, 58, 6955 (1998).
408. G.V. Osipov and J.J. Collins, Phys. Rev. E, 60, 54 (1999).
409. M. Allessie, C. Kirchhof, G.J. Scheffer, F. Chorro, and J. Brugada, Circulation,
84, 1689 (1991).
410. A. Capucci, F. Ravelli, G. Nollo, A.S. Montenero, M. Biffi, and G.Q. Villani,
J. Cardiovasc. Electrophysiol., 10 319 (1999).
411. E.G. Daoud, B. Pariseau, M. Niebauer, F. Bogun, R. Goyal, M. Harvey,
K.C. Man KC, S.A. Strickberger, and F. Morady, Circulation, 94, 1036 (1996).
412. J.M. Kalman, J.E. Olgin, M.R. Karch, and M.D. Lesh, J. Cardiovasc. Elec-
trophysiol., 7, 867 (1996).
413. C. Kirchhof, F. Chorro, G.J. Scheffer, J. Brugada, K. Konings, Z. Zetelaki,
and M. Allessie, Circulation, 88, 736 (1993).
414. B.H. KenKnight, P.V. Bayly, R.J. Gerstle, D.L. Rollins, P.D. Wolf, W.M.
Smith, and R.E. Ideker, Circ. Res., 77, 849 (1995).
415. K.J. Lee, Phys. Rev. Lett. 79, 2907 (1997).
416. R. Mantel and D. Barkley, Phys. Rev. E, 54, 4791 (1996).
417. A.T. Winfree, Physica D, 49, 125 (1991).
418. F. Xie, Z. Qu, J.N. Weiss, and A. Garfinkel, Phys. Rev. E, 59, 2203 (1999).
419. G. Gottwald, A. Pumir, and V. Krinsky, Chaos, 11, 487 (2001).
420. A.L. Bassett, S. Chakko, and M. Epstein, J. Hypertens., 15, 915 (1997).
421. T.R. Chay, J. Electrocardiol., 28, 191 (1995).
422. V.I. Krinsky and K.I. Agladze, Physica D, 8, 50 (1983).
423. V.S. Zykov, Ann. N.Y. Acad. Sci., 591, 75 (1990).
424. A.T. Winfree, Ann. N.Y. Acad. Sci., 591, 190 (1990).
425. J.J. Tyson and J.P. Keener, Physica D, 32, 327 (1988).
426. A.N. Zaikin and A.M. Zhabotinsky, Nature, 225, 535 (1970).
427. M.R. Guevara, A. Shrier, and L. Glass, Am. J. Physiol., 254, H1 (1988).
428. T.J. Lewis and M.R. Guevara J. Theor. Biol., 146, 407 (1990).
429. Z. Qu and A. Garfinkel, IEEE Trans. Biomed. Eng., 46, 1166 (1999).
430. S. Rush and H. Larsen, IEEE Trans. Biomed. Eng., 25, 389 (1978).
431. Z. Qu, J.N. Weiss, and A. Garfinkel, Phys. Rev. E 61, 727 (2000).
432. P.Q. Anderson, J.E. Knoben, and W.G. Troutman, in Handbook of Clinical
Drug Data. 9 th ed. Stamford, CT: Appleton & Lange; 345 (1999).
433. S.C. Vlay, in A Practical Approach to Cardiac Arrhythmias. 2nd ed. Boston,
MA: Little, Brown & Company; 309 (1996).
434. N. Marwan, N. Wessel, U. Meyerfeldt, A. Schirdewan, and J. Kurths, Phys.
Rev. E 66, 026702 (2002).
435. L. Gammaitoni, P. Hanggi, P. Jung, and F. Marchesoni, Rev. Mod. Phys. 70,
223 (1998).
436. A. Neiman, A. Silchenko, V. Anishchenko, and L. Schimansky-Geier, Phys.
Rev. E 58, 7118 (1998).
437. Z.F. Mainen and T.J. Sejnowski, Science 268, 1503 (1995).
438. O.N. Bjornstad, R.A. Ims, and X. Lambin, Trends Ecol. Evol. 14, 427 (1999).
439. P.J. Hudson and I.M. Cattadori, Trends Ecol. Evol. 14, 1 (1999).
358 References
477. M. Zaks, E.H. Park, M. Rosenblum, and J. Kurths, Phys. Rev. Lett. 82, 4228
(1999).
478. C.S. Zhou, J. Kurths, I.Z. Kiss, and J.L. Hudson, Phys. Rev. Lett. 89, 014101
(2002).
479. E. Ott, Chaos in Dynamical Systems (University Press, Cambridge, 1993).
480. I.Z. Kiss and J.L. Hudson, Phys. Rev. E 64, 046215 (2001).
481. I.Z. Kiss, Y. Zhai, J.L. Hudson, C.S. Zhou and J. Kurths, Chaos, 13, 267
(2003).
482. B. Lindner, J. Garcia-Ojalvo, A. Neiman, and L. Schimansky-Geier 2004 Phys.
Rep. 392 321–424.
483. A.S. Pikovsky and J. Kurths, Phys. Rev. Lett. 78, 775 (1997).
484. B. Hu and C.S. Zhou, Phys. Rev. E 63, 026201 (2001).
485. C.S. Zhou, J. Kurths, and B. Hu, Phys. Rev. Lett. 87, 098101 (2001).
486. A. Neiman, L. Schimansky-Geier, A. Cornell-Bell, and F. Moss, Phys. Rev.
Lett. 83, 4896 (1999).
487. J. Balakrishnan, Phys. Rev. E 73, 036206 (2006).
488. C.S. Zhou, J. Kurths, and B. Hu, Phys. Rev. E 67, 030101(R) (2003).
489. V. Petrov, Q. Ouyang, and H.L. Swinney, Nature (London) 388, 655 (1997).
490. A.L. Lin, M. Bertram, K. Martinez, and H.L. Swinney, A. Ardelea and
G.F. Carey, Phys. Rev. Lett. 84, 4240 (2000).
491. V.K. Vanag, L. Yang, M. Dolnik, A.M. Zhabotinsky, and I.R. Epstein, Nature
(London) 406, 389 (2000).
492. M. Kim, M. Bertram, M. Pollmann, A. von Oertzen, A.S. Mikhailov, H.H.
Rotermund, and G. Ertl, Science 292, 1357 (2001).
493. C.S. Zhou and J. Kurths, Phys. Rev. E 69, 056210 (2004).
494. C.S. Zhou and J. Kurths, New J. Phys. 7, 18 (2005).
495. R. Albert and A.-L. Barabási, Rev. Mod. Phys. 74, 47 (2002).
496. P. Erdös and A. Rényi, Publ. Math. Inst. Hung. Acad. Sci. 5, 17 (1959).
497. D.J. Watts and S.H. Strogatz, Nature (London) 393, 440 (1998).
498. A.-L. Barabási and R. Albert, Science 286, 509 (1999).
499. H. Jeong, B. Tombor, R. Albert, Z.N. Oltavi, and A.-L. Barabási, Nature 407,
651 (2000).
500. M.E.J. Newman, Proc. Natl Acad. Sci. USA 98, 404 (2001).
501. L.A.N. Amaral, A. Scala, M. Barthélémy, and H.E. Stanley, Proc. Natl Acad.
Sci. USA 97, 11149 (2000).
502. O. Sporns, G. Tononi, and G.M. Edelman, Cereb. Cortex 10, 127 (2000).
503. D.J. Watts, Small Worlds (Princeton University Press, Princeton, 1999).
504. S.H. Strogatz, Nature (London) 410, 268 (2001).
505. S.N. Dorogovtsev and J.F.F. Mendes, Adv. Phys. 51, 1079 (2002).
506. M.E.J. Newman, SIAM Rev. 45, 167 (2003).
507. R. Albert, H. Jeong, and A.-L. Barabási, Nature (London) 406, 378 (2000).
508. D.S. Callaway, M.E.J. Newman, S.H. Strogatz, and D.J. Watts, Phys. Rev.
Lett. 85, 5468 (2000).
509. R. Chen, K. Erez, D. ben Avraham, and S. Havlin, Phys. Rev. Lett. 85, 4626
(2000).
510. R. Pastor-Satorras and A. Vespignani, Phys. Rev. Lett. 86, 3200 (2001).
511. R.M. May, and A.L. Lloyd, Phys. Rev. E 64, 066112 (2001).
512. M.E.J. Newman, Phys. Rev. E 64, 016128 (2002).
513. A.E. Motter, Phys. Rev. Lett. 93, 098701 (2004).
360 References
514. S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, and D.-U. Hwang, Phys. Rep.
424, 175 (2006).
515. L.F. Lago-Fernández, R. Huerta, F. Corbacho, and J.A. Sigüenza, Phys. Rev.
Lett. 84, 2758 (2000).
516. N. Masuda and K. Aihara, Biol. Cybern. 90, 302 (2004).
517. X. Guardiola, A. Diaz-Guilera , M. Llas, and C. J. Perez, Phys. Rev. E 62,
5565 (2000).
518. H. Hasegawa, Phys. Rev. E 70, 066107 (2004).
519. H. Hong, M. Y. Choi, and B.J. Kim, Phys. Rev. E 65, 026139 (2002).
520. A.M. Batista, S.E.D. Pinto, R.L. Viana, and S.R. Lopes, Physica A 322, 118
(2003).
521. P.M. Gade and C.K. Hu, Phys. Rev. E 62, 6409 (2000).
522. J. Jost and M.P. Joy, Phys. Rev. E 65, 016201 (2001).
523. M. Barahona and L.M. Pecora, Phys. Rev. Lett. 89, 054101 (2002).
524. S. Sinha, Phys. Rev. E 66, 016209 (2002).
525. I. Belykh, V. Belykh, and M. Hasler, Physica D 195, 159 (2004).
526. T. Nishikawa, A.E. Motter, Y.-C. Lai, and F.C. Hoppensteadt, Phys. Rev.
Lett. 91, 014101 (2003).
527. F. Chung and L. Lu, Proc. Natl. Acad. Sci. U.S.A. 99, 15879 (2002).
528. R. Cohen and S. Havlin, Phys. Rev. Lett. 90, 058701 (2003).
529. L.M. Pecora and T.L. Carroll, Phys. Rev. Lett. 80, 2109 (1998).
530. M.E.J. Newman, C. Moore, and D.J. Watts, Phys. Rev. Lett. 84, 3201 (2000).
531. M.E.J. Newman, S.H. Strogatz, and D.J. Watts, Phys. Rev. E 64, 026118
(2001).
532. S. De Monte, F. d’Ovidio, and E. Mosekilde, Phys. Rev. Lett. 90, 054102
(2003).
533. S. Jalan and R.E. Amritkar, Phys. Rev. Lett. 90, 014101 (2003).
534. S. Jalan and R.E. Amritkar, Physica A 321, 220 (2003).
535. S. Jalan and R.E. Amritkar, Physica A 346, 13 (2005).
536. L.G. Morelli, H.A. Cerdeira, and D.H. Zanette, Eur. Phys. J. B 43, 243 (2005)
537. I. Stewart, Nature (London) 427, 601 (2004).
538. C.S. Zhou and J. Kurths, Chaos 16, 015104 (2006).
539. A. Barrat, M. Barthélemy, R. Pastor-Satorras, and A. Vespignani, Proc. Natl.
Acad. Sci. U.S.A. 101, 3747 (2004).
540. S.H. Yook, H. Jeong, A.-L. Barabási, and Y. Tu, Phys. Rev. Lett. 86, 5835
(2001).
541. M.E.J. Newman, Phys. Rev. E 64, 016132 (2001).
542. L.A. Braunstein, S.V. Buldyrev, R. Cohen, S. Havlin, and H.E. Stanley, Phys.
Rev. Lett. 91, 168701 (2003).
543. D.J. Felleman and D.C. Van Essen, Cerebral Cortex 1, 1 (1991).
544. J.W. Scannell, G.A.P.C. Burns, C.C. Hilgetag, M.A. O’eil, and M.P. Yong,
Cerebral Cortex 9, 277 (1999).
545. C.J. Stam and E.A. de Bruin, Hum. Brain Mapp. 22, 97 (2004).
546. E. Salinas and T.J. Sejnowski, Nature Neurosci. 2, 539 (2001).
547. P. Friés, Trends Cogn. Sci. 9, 474 (2005).
548. A. Schnitzler and J. Gross, Nature Neurosci. 6, 285 (2005).
549. O. Sporns, D.R. Chialvo, M. Kaiser, and C.C. Hilgetag, Trends Cogn. Sci. 8,
418 (2004).
550. C.J. Stam, Neurosci. Lett. 355, 25 (2004).
References 361
551. V.M. Eguı́luz, D.R. Chialvo, G. Cecchi, M. Baliki, and A.V. Apkarian, Phys.
Rev. Lett. 94, 018102 (2005).
552. R. Salvador et al., Cereb. Cortex 15, 1332 (2005).
553. C.S. Zhou, L. Zemanová, G. Zamora, C.C. Hilgetag, and J. Kurths, Phys. Rev.
Lett. 97, 238103 (2006); New J. Phys. (2007) (to appear).
554. L. Zemanová, C.S. Zhou, and J. Kurths, Physica D 224, 202 (2006).
555. B.T. Grenfell, O.N. Bjornstad, and J. Kappey, Nature (London) 414, 716
(2001).
556. N.C. Grassly, C. Fraser, and G.P. Garnett, Nature( London) 433, 417 (2005).
557. J.G. Restrepo, E. Ott, and B.R. Hunt, Phys. Rev. Lett. 93, 114101 (2004).
558. M. Denker, M. Timme, M. Diesmann, F. Wolf, and T. Geisel, Phys. Rev. Lett.
92, 074103 (2004).
559. X.F. Wang, Int. J. Bifurcation Chaos Appl. Sci. Eng. 12, 885 (2002).
560. A.E. Motter, C.S. Zhou, and J. Kurths, Europhys. Lett. 69, 334 (2005).
561. A.E. Motter, C.S. Zhou, and J. Kurths, Phys. Rev. E 71, 016116 (2005).
562. S.N. Dorogovtsev and J.F.F. Mendes, Phys. Rev. E. 62, 1842 (2000).
563. F.R.K. Chung, Spectral Graph Theory (AMS, Providence, 1994).
564. F. Chung, L. Lu, and V. Vu, Proc. Natl. Acad. Sci. U.S.A. 100, 6313 (2003).
565. C.S. Zhou, A.E. Motter, and J. Kurths, Phys. Rev. Lett. 96, 034101 (2006).
566. G. Korniss, M.A. Novotny, H. Guclu, Z. Toroczkai, and P.A. Rikvold, Science
299, 677 (2003).
567. M. Chavez, D.-U. Hwang, A. Amann, H.G.E. Hentschel, and S. Boccaletti,
Phys. Rev. Lett. 94, 218701 (2005).
568. D.-U. Hwang, M. Chavez, A. Amann, and S. Boccaletti, Phys. Rev. Lett. 94,
138701 (2005).
569. J. Lu, X. Yu and G. Chen, Physica A 334, 281 (2004).
570. I.V. Belykh, V.N. Belykh, and M. Hasler, Physica D 195, 188 (2004).
571. R.E. Amritkar and C. Hu, Chaos, 16, 015117 (2006).
572. C.S. Zhou and J. Kurths, Phys. Rev. Lett. 96, 164102 (2006).
573. T. Gross, C.J. Dommar D’Lima, and B. Blasius, Phys. Rev. Lett. 96, 208701
(2006).
Index
Complete synchronization, 4, 68, 70, 72, two, 15, 31, 90, 110, 111, 114, 119,
76, 80, 85, 86, 269 152, 179, 188, 249, 251–253, 255,
Convection, 18, 19, 169 257, 261, 262, 265, 270, 278, 279
Correlation Degree, 16, 25, 70, 72, 117, 121, 124,
cross, 72 140, 147, 163
degree, 333, 337 Distribution
increasing, 295 contraction region, 275
noise, 293 coupling
strong, 70, 72, 73, 84 local, 226, 228, 318
Coupled global, 56, 124, 179, 230, 231, 295,
map, 77, 91, 169, 170, 172, 185, 189, 297, 298, 301, 318
190, 192, 195, 197, 211, 333 of degree, 319, 328, 334, 337
unidirectionaly, 90, 103, 128, 219, 220 gaussian, 124, 157, 269, 271
Coupling of phase, 133, 141, 176
asymmetric, 189, 195, 197, 210, 239,
240, 320, 333 E
diffusive, 131, 150, 152, 171, 189, 207,
Embedding, 174, 178
216, 221, 222, 305, 306, 314
Energy, 252, 271, 327
dissipative, 58, 61, 62, 235, 241
Ensemble average, 17
global, 6, 55, 103, 123–125, 213, 225,
Entrainment, 5, 8, 38, 63, 68, 84, 90, 96,
230, 231, 295–299, 301, 317, 324,
123, 171, 180, 200, 230
325, 327
Excitability, 305, 308
mean field, 179, 180–185, 295
Excitable system, 4, 8, 11, 29–34, 36,
nearest neighbor, 6, 125, 131, 225,
52, 269, 312
313, 317, 321
nonlinear, 188, 189, 213
reactive, 234, 241, 242, 244 F
scalar, 152 Fast motion, 12, 31, 172
symmetric, 90, 112, 113, 115, 117, Feedback, 213–216, 218–220, 222, 224,
124, 189, 190, 195–198, 239, 240, 225, 228, 230, 231, 281, 282, 284,
241, 320, 333, 337 314
unidirectional, 106, 110, 213, 214, Filter, 24, 215, 224
218, 234 Firefly, 22
Coupling function, 57, 58, 105, 114 Focus, 5, 7, 9, 38, 48, 59, 105, 170, 178,
Crosscorrelation, 72 198, 200, 235, 278
Fokker–Planck equation, 271
D Force
Data, 34, 90, 137, 140, 166, 259 periodic, 4, 35, 36, 40, 47, 52, 130,
Decoherence, 167 179, 251, 270, 288, 289
Demodulation, 4 random, 269, 270, 273
Detuning, 68 Fourier, 152
Differential equation, 18, 19, 28, 29, 48, Fractal, 178, 185
170, 234, 236 Frequency
Diffusion, 76, 152, 240 instantaneous, 14, 69, 121
constant, 16, 17, 41, 70, 81–84, 87, mean, 16, 18, 19, 28, 35, 36, 40,
253, 309 43, 61, 63, 69, 74, 80–83, 85–88,
Dimension 113, 116, 119, 120, 131, 133, 153,
one, 7, 104, 110, 243, 247, 251–253, 155, 161, 217, 218, 220, 221, 223,
255, 256, 258–262, 264, 274, 278 227–229, 272, 296–298
Index 365
Population, 3, 9, 123, 169, 189, 207, Signal-to-noise ratio, 303, 306, 307
270, 301, 340 Slow motion, 31
Potential, 29–31, 33, 34, 53, 230, 251, Small world, 317–319, 321
254–257, 271, 279, 297, 300, 340 Soft transition, 63, 87, 115–117, 122,
Power spectrum, 12, 68, 114, 301, 302 123, 128, 141, 142, 155, 156, 167,
Pulse, 31, 34, 52, 53, 126, 177, 187, 171, 203, 211, 221, 227
252–259, 261, 265, 305, 306, 308, Spatiotemporal chaos, 177, 178, 187
319, 332 spectrum
Lyapunov, 84, 89, 117, 155, 157
Q eigenvalues, 333, 337
Quasiperiodic, 59, 109–112, 114, 179, Spike
253, 264, 266 train, 29, 30, 278, 280, 281, 311, 312
Spiral wave, 34, 53, 177, 251–258,
R 260–266
Random walk, 242 Splay state, 193
Stability, 7, 59, 80, 91, 113, 152,
Relaxation oscillator, 56, 149
189–191, 193–196, 198, 204,
Repeller, 45, 46
209, 242, 257, 279, 319–321,
Resonance 333, 334
coherent, 305 Stimulus, 33, 52, 53, 251, 252, 254–264,
parametric, 48 269
stochastic, 249, 269 Stroboscopic technique, 40
array-enhanced, 313 Suppression of oscillations, 160
Return time Symmetry, 115, 117, 140, 147, 242, 331
average, 41, 45, 294 Synchronization
Resonance-like behaviour, 307 Cluster, 114, 116, 118, 120, 128, 130,
Rhythm, 3, 5, 7, 8, 22, 28, 53, 179, 180, 135, 145, 146, 149, 150, 181, 204,
182, 252 207, 227, 302, 326, 339, 340
Riddling, 47 external, 4, 35, 37, 39, 41, 47, 52
Rotation number, 25, 26, 45, 90–93, 95, global, 8, 105, 106, 108, 113–116,
99, 173, 200, 203–208, 211 118, 123, 126–129, 133–136,
Rotator, 21, 22, 24, 55, 56, 66, 111, 113, 140, 144
119, 120, 123, 126, 127, 206, 233 mutual, 4, 28, 55, 77, 103, 114, 123,
Roessler system, 166 130, 155, 180, 187, 238
Synchronization region, 40, 42, 45, 49,
50, 61, 64, 67, 72, 74, 75, 84, 88,
S
91, 92, 95, 97, 183–185, 190–192,
Saddle, 23, 24, 38, 42, 46, 47, 63, 74, 196, 197, 204, 210, 271, 272, 290,
110, 113, 278–282, 284, 289–291 291, 293
Saddlenode, 23 Synchronization transition, 7, 41, 73,
Scale free, 318, 319, 324, 325, 327 99, 117, 119, 120, 122, 124, 129,
Self-sustained oscillator, 12, 36, 305, 167, 185, 222, 228
313, 314 Synchrony, 8, 53, 125, 251, 252, 254–261,
Shannon entropy, 287 270
Signal System
chaotic, 45, 89, 90 autonomous, 51, 77
periodic, 35, 39, 178, 179, 218, 291, bistable, 249
292, 310, 311, 313–315 distributed, 127
368 Index