1: \documentclass[aps,prl,twocolumn,showpacs,floatfix]{revtex4}
2: \usepackage{graphics}
3: \usepackage{epsfig}
4: \usepackage{subfigure}
5: \usepackage{amsmath,amsfonts,amssymb,graphicx}
6:
7: \begin{document}
8: \title{Synchronization reveals correlation between oscillators on networks}
9:
10: \author{Jie Ren$^{1}$}
11: \author{Huijie Yang$^{2}$}
12: %\author{Yi-Cheng Zhang$^{1}$}
13: \affiliation{$^{1}$Department of Physics, University of Fribourg,
14: CH-1700, Fribourg, Switzerland\\
15: $^{2}$ Department of Physics and Center for Computational Science
16: and Engineering, National University of Singapore, Singapore 117542}
17:
18: \date{today}
19:
20: \begin{abstract}
21: The understanding of synchronization ranging from natural to social
22: systems has driven the interests of scientists from different
23: disciplines. Here, we have investigated the synchronization dynamics
24: of the Kuramoto dynamics departing from the fully synchronized
25: regime. We have got the analytic expression of the dynamical
26: correlation between pairs of oscillators that reveals the relation
27: between the network dynamics and the underlying topology. Moreover,
28: it also reveals the internal structure of networks that can be used
29: as a new algorithm to detect community structures. Further, we have
30: proposed a new measure about the synchronization in complex networks
31: and scrutinize it in small-world and scale-free networks. Our
32: results indicate that the more heterogeneous and ``smaller" the
33: network is, the more closely it would be synchronized by the
34: collective dynamics.
35: \end{abstract}
36:
37: \pacs{05.45.Xt, 89.75.Fb}
38:
39: \maketitle
40:
41: The last decade has witnessed the rapid explosion of interest in
42: complex networks, which are found in many fields as diverse as
43: biology, technology and social organizations
44: \cite{review1,review2,review3,review4}. The relation between
45: structure and function becomes a key area in the study of complex
46: networks. In the study of networked-dynamics, the emergent
47: synchronization of interacting oscillators generally has occupied a
48: privileged position because of its rich applications in variety of
49: areas ranging from Neuroscience to Sociology
50: \cite{sync1,sync2,sync3,sync4,sync5,sync6,sync7,sync8,sync9,sync10,sync11,sync12,sync13}.
51: In neural systems, the brain dynamics is characterized by
52: synchronization phenomena for a given topology of synapses. And in
53: metabolic networks, hundreds of interconnected biochemical reactions
54: are responsible for the biomass and energy fluxes which are adjusted
55: to optimize the robustness of synchronized behavior. Other example
56: is the synchronized or coordinated behavior of communication
57: patterns in social organizations. The study of synchronization
58: provides us with insights into the key issue: how the collective
59: dynamics couple the relationships between the systematic function
60: and the underlying topology.
61:
62: One of the most successful attempts to understand synchronization
63: phenomena is due to the Kuramoto model (KM) \cite{KM,KM1}, which is
64: rich enough for many different contexts, including superconducting
65: currents in Josephson junction arrays, emerging coherence in
66: populations of chemical oscillators, and the accuracy of central
67: circadian pacemakers in insects and vertebrates, \emph{etc.}. This
68: model describes a population of $N$ coupled phase oscillators which
69: evolves in time according to the following dynamics
70: \begin{equation}
71: \frac{d\theta_i}{dt} = \omega_i + \sigma
72: \sum_{j}A_{ij}\sin(\theta_{j}-\theta_{i}) + \xi_{i}(t).\label{eq:KM}
73: \end{equation}
74: Here, $\theta_i$ represents the phase of the $i$th oscillator,
75: $\omega_i$ the intrinsic frequency, $\sigma$ the coupling constant,
76: $A_{ij}$ the effective coupling between the oscillators, and
77: $\xi_{i}(t)$ is the white noise due to the complicated environment,
78: with expectation and variance
79: \begin{eqnarray}
80: \langle \xi_i(t)\rangle &=& 0, \nonumber \\
81: \langle \xi_i(t) \xi_j(t') \rangle &=& 2 \delta_{ij} \delta(t-t').
82: \nonumber
83: \end{eqnarray}
84: There are many attractors that each oscillator rotates at its
85: natural frequency when incoherent. As the coupling strength
86: increases, some units become resonant. At sufficiently strong
87: coupling, when coherent, there is only one attractor of the
88: dynamics, and all oscillators rotate at the same frequency,
89: $\overline\omega_i$, the average of $\omega_i$. We call it fully
90: synchronized.
91:
92: The synchronization problem is solved mainly in the mean-field
93: approach which, unfortunately, is not usually fulfilled in real
94: systems. The nontrivial pattern of connectivity in complex networks
95: brings the non mean-field properties and incorporates many new
96: questions to the research of synchronization. Because of the elegant
97: work of Pecora and Carroll \cite{MSF}, the Master Stability Function
98: (MSF) formalism allows us to analysis the dynamical
99: synchronizability in terms of purely topological porperties,
100: independent of details of unit dynamics. However, the MSF requires
101: several constraints and has limited applications that it only deals
102: with the stability of the exact synchronized states, \emph{i.e.}, it
103: is used to study infinitesimal deviations from the synchronization
104: manifold of chaotic oscillators. However, many realistic dynamics
105: are only close to or even far from the full synchronization.
106:
107: In this work we study the KM dynamics departing from the fully
108: synchronized state. We get the analytic expression of the dynamical
109: correlation between any two oscillators which reveals the relation
110: between the network dynamics and the underlying topology. It could
111: be used as a new algorithm to detect community structures. Moreover,
112: we propose a new measure about the synchronization in complex
113: networks and scrutinize it in Small-World and Scale-Free networks.
114:
115: \begin{figure}
116: \scalebox{0.75}[0.9]{\includegraphics{net.eps}}
117: \scalebox{0.85}[0.9]{\includegraphics{cor.eps}}
118: \caption{\label{fig:SFI} Top row: Illustration of the Santa Fe
119: Institute collaboration network. Bottom left: The analytic
120: correlation matrix resulting from Eq. \ref{eq:cor}. Bottom right:
121: The simulated results of the correlation between pairs of
122: oscillators from the Kuramoto model. The colors are a gradation
123: between white (strong correlation) and black (weak correlation).}
124: \end{figure}
125:
126: When the KM dynamics is close to the attractor of synchronization,
127: the phase differences are small and then the sine coupling function
128: can be approximated linearly, so that we can analyze the linearized
129: dynamics of the Kuramoto model in terms of the Laplacian matrix,
130: \begin{eqnarray}
131: \frac{d\theta_i}{dt} &=& - \sigma \sum_{j}a_{ij}(\theta_{i}-\theta_{j}) + \xi_{i}(t) \nonumber \\
132: &=& - \sigma
133: \sum_{j}L_{ij}\theta_{j} + \xi_{i}(t), \label{eq:Laplace}
134: \end{eqnarray}
135: where
136: $L_{ij}=\delta_{ij}\sum_{m}a_{im}-a_{ij}=\delta_{ij}k_i-a_{ij}$.
137: $a_{ij}$ stands for the adjacency matrix ($1$ if nodes $i$ and $j$
138: are connected and $0$ otherwise), $\delta_{ij}$ is the Kronecker
139: delta and the degree $k_i$ is defined as the number of nodes to
140: which the node $i$ is connected. Thus, we can rewrite the equation
141: in terms of the normal modes,
142: \begin{equation}
143: \frac{d\vartheta_{\alpha}}{dt} = - \sigma
144: \lambda_{\alpha}\vartheta_{\alpha} + \zeta_{\alpha}(t),
145: \end{equation}
146: where $\vartheta_{\alpha} = \sum_{j}\psi_{\alpha j}\theta_{j}$,
147: $\zeta_{\alpha} = \sum_{j}\psi_{\alpha j}\xi_{j}$, and $\psi_{\alpha
148: j}$ denotes the $\alpha$th normalized eigenvector of the Laplacian,
149: $\lambda_{\alpha}$ is the corresponding eigenvalue for
150: $\alpha=0,...,N-1$. Considering $\xi_{i}(t)$ is delta correlated, we
151: can find easily that $\zeta_{\alpha}(t)$ is also delta correlated:
152: $\langle\zeta_{\alpha}(t)\zeta_{\beta}(t')\rangle=2\delta_{\alpha\beta}\delta(t-t')$.
153:
154: Without loss of generalization, we set $\sigma=1$,
155: $\zeta_{\alpha}(0)=0$. Using stochastic calculus methods
156: \cite{method}, we get
157: \begin{equation}
158: \vartheta_{\alpha}(t) = \int_{0}^t e^{- \lambda_{\alpha} (t-t')}
159: \zeta_{\alpha}(t') dt',
160: \end{equation}
161: so that
162: \begin{eqnarray}
163: \langle \vartheta_{\alpha}(t)^2 \rangle &=& \int_{0}^t \int_{0}^t
164: e^{- \lambda_{\alpha} (t-t')} e^{- \lambda_{\alpha} (t-t'')}\langle \zeta_{\alpha}(t') \zeta_{\alpha}(t'') \rangle dt' dt'' \nonumber\\
165: &=& \int_{0}^t 2 e^{-2 \lambda_{\alpha} (t-t')} dt' \nonumber\\
166: &=& \frac{1}{ \lambda_{\alpha}}(1-e^{-2 \lambda_{\alpha}t})
167: \end{eqnarray}
168: Thus, the time scales of the dynamical relaxation of the eigenmodes
169: are inversely proportional to the corresponding eigenvalues. And for
170: large $t\gg 1/2\lambda_{\alpha}$, we have
171: \begin{equation}
172: \langle \vartheta_{\alpha}(t)^2 \rangle=1/\lambda_{\alpha}.
173: \end{equation}
174: Transforming it back to the basis of $\theta_{i}$ by substituting
175: $\theta_{i}=\sum_{\alpha}\psi_{\alpha i}\vartheta_{\alpha}$, we find
176: the steady-state correlation function:
177: \begin{eqnarray}
178: \langle (\theta_i-\overline \theta) (\theta_j-\overline
179: \theta)\rangle &=& \sum_{\alpha=1}^{N-1} \psi_{\alpha i}
180: \psi_{\alpha j}\langle \vartheta_{\alpha}^2 \rangle \nonumber\\
181: &=& \sum_{\alpha=1}^{N-1} \frac{1}{\lambda_{\alpha}} \psi_{\alpha i}
182: \psi_{\alpha j}, \label{eq:cor}
183: \end{eqnarray}
184: where $\overline \theta = \frac{1}{N}\sum_i \theta_i$ and
185: $\langle...\rangle$ stands for the average over initial random
186: phases. This analytic expression gives the dynamical correlation
187: between any pair of oscillators just in terms of the properties of
188: the underlying topology. Actually, the right term of Eq.
189: \ref{eq:cor} gives the pseudo-inverse of the Laplacian matrix, which
190: acts like a Green Function. It could be used to detect the community
191: structure in various complex networks.
192:
193: One real world example with known community structure is the
194: collaboration network of the Santa Fe Institute from ref.
195: \cite{Girvan}. The top row of Fig. 1 illustrates the largest
196: component of the collaboration graph, which consists of $118$
197: scientists denoted by nodes. The edge is drawn between a pair of
198: scientists if they coauthored one or more papers. Obviously, the
199: scientists of different interests group in different communities.
200:
201: In the bottom row of Fig. 1, we illustrate the dynamical correlation
202: between any pairs of nodes on which the KM oscillators is placed.
203: The left one is the analytic results from the relation Eq.
204: \ref{eq:cor} and the right is the results of numerical simulations
205: from Eq. \ref{eq:KM}. They accord with each other very well. In the
206: two figures we can identify the distinct dynamics-based communities
207: wherein nodes become strong correlated in groups, coherently with
208: their topological structure. Moreover, smaller subcommunities can be
209: found clearly in these dynamics-based communities which corresponds
210: with the sketch map of the collaboration network. Although some
211: nodes in the same dynamics-based community are not directly
212: connected, they are topologically equivalent and belong to the same
213: structural level. Therefore, they have strong correlation on
214: dynamics. The dynamics-based groups indicates the different
215: functional modules. When applied to the specific internal structure,
216: which might be the fingerprint of different functional groups for
217: social or biological networks, the dynamical correlation may help us
218: understand better the interplay between systemic structure and
219: function.
220:
221: \begin{figure}
222: \scalebox{0.8}[0.80]{\includegraphics{k_lamda.eps}}
223: \caption{\label{fig:epsart} (color online). Illustration of the same
224: hierarchical structure in the eigenvalues spectra and degree series.
225: It implies a progressive cascade of the synchronization pattern from
226: the central core to the nodes with smaller degrees. The network data
227: is used as the same as in Fig. \ref{fig:SFI}.}
228: \end{figure}
229:
230: %=============Also, we can get
231: %\begin{eqnarray}
232: %\langle (\theta_i-\theta_j)^2 \rangle &=& \sum_{\alpha=1}^{N-1}
233: %(\psi_{\alpha i} - \psi_{\alpha j})^2 \langle \vartheta_{\alpha}^2\rangle \nonumber\\
234: %&=& \sum_{\alpha=1}^{N-1} \frac{1}{\lambda_{\alpha}} (\psi_{\alpha
235: %i} - \psi_{\alpha j})^2.
236: %\end{eqnarray}
237: %===========================(This part is not necessary, may be moved to the reference?)
238:
239: In another way, we can use the mean-field approximation to rewrite
240: the Eq. \ref{eq:Laplace} as:
241: \begin{eqnarray}
242: \frac{d\theta_i}{dt} &=& - \sum_{j}A_{ij}(\theta_{i} - \theta_{j}) +
243: \xi_{i}(t) \nonumber\\
244: &=& - k_{i}(\theta_{i} - \overline \theta) + \xi_{i}(t).
245: \end{eqnarray}
246: Using the same stochastic calculus method, we get
247: \begin{eqnarray}
248: \langle (\theta_i-\overline \theta)^2\rangle = \frac{1}{k_i}(1-e^{-2
249: k_i t}),
250: \end{eqnarray}
251: which indicates that the time scales of the dynamical relaxation of
252: the observed phases are inversely proportional to the degrees of the
253: corresponding oscillators. The different connectivities in the
254: topology give rise to the corresponding order of the eigenvalues,
255: which are illustrated in Fig. 2. The hubs with the large eigenvalues
256: induce the fast relaxations, \emph{i.e.}, the evolutional pattern of
257: synchronization starts first at the hubs and then pervades almost
258: the whole network in a hierarchical structure across the nodes with
259: smaller degrees. And for large $t\gg1/2k_i$, we have got $\langle
260: (\theta_i-\overline \theta_i)^2\rangle = 1/k_i$, which indicates
261: that hubs are synchronized more closely by the collective dynamics.
262:
263: For a given realization of networks, we use $R$ to denote the
264: average variance of phase variables which characterizes the
265: collective synchronized ability of the underlying network:
266: \begin{eqnarray}
267: R = \frac{1}{N} \sum^{N}_{i}\langle
268: (\theta_i-\overline\theta)^2\rangle ,
269: \end{eqnarray}
270: The smaller the value of $R$, the smaller the fluctuation of network
271: is, \emph{i.e.}, the more synchronized . Substituting the Eq.
272: \ref{eq:cor} and considering the orthonormality of $\psi_{\alpha
273: i}$, we have
274: \begin{eqnarray}
275: R = \frac{1}{N} \sum_{\alpha=1}^{N-1} \frac{1}{\lambda_{\alpha}}.
276: \label{eq:sync}
277: \end{eqnarray}
278: Therefore, independent of details of unit dynamics, we can
279: characterize the dynamical synchronizability of networked systems
280: just in terms of the spectra of the Laplacian matrix, \emph{i.e.},
281: the purely topological porperties. It is the same as the spirit of
282: the MSF. However, in contrast to the MSF, the new measure Eq.
283: \ref{eq:sync} considers the spectrum in its entirety, not only the
284: extremal eigenvalues. What this new measure emphasizes is on the
285: fluctuation of the process of synchronization, rather than on
286: rigorous bounds for the threshold of desynchronization. In the
287: following, we use this new quantity about synchronization, $R$ to
288: measure the synchronizable ability in small-world (SW) and
289: scale-free (SF) networks, respectively.
290:
291: \begin{figure}
292: \scalebox{0.8}[0.85]{\includegraphics{SW_SF.eps}}
293: \caption{\label{fig:swsf} (a) The synchronizability of SW networks
294: of size $N=1000$, $\langle k \rangle=8$. The average network
295: distance $D$ and the new measure $R$ decreases with the rewiring
296: probability $p$ in the same trend, while the standard deviation $S$
297: of the degree distribution increases. (b) The synchronizability of
298: SF networks of size $N=1000$, $k_0=4$. The values of $D$ and $R$
299: increase with the scaling exponent $\gamma$, while the values of $S$
300: are decreased. Since less $R$ means more synchronizable, all the
301: results in SW and SF networks indicate that the network would be
302: more synchronizable as it becomes ``smaller" and more heterogeneous.
303: All plots are averaged over 100 realizations.}
304: \end{figure}
305:
306: We first consider the Watts-Strogatz model of the SW networks
307: \cite{WS}, which is constructed by a rewiring process on a regular
308: ring graph. The probability of rewiring connection is controlled by
309: a parameter $p$, by tuning which the obtained network possesses both
310: short average distance and high heterogeneity that nodes no longer
311: have the same degrees, instead they follow a Poisson distribution.
312: We define $S$ as the standard deviation of the degree distribution
313: and $D$ as the average network distance, averaged over all pairs of
314: nodes. The upper layer of Fig. \ref{fig:swsf}(a) shows the
315: dependence of $S$ and $D$ on the rewiring probability $p$ in the SW
316: network. As $p$ increases, the variance $S$ increases, which implies
317: that the degree distribution becomes more broad and heterogeneous.
318: And as expect, the network distance $D$ decreases as the rewiring
319: probability $p$ (or the heterogeneity of the degree distribution)
320: increases. In the lower layer of Fig. \ref{fig:swsf}(a), $R$ is
321: observed to decrease, which implies enhancement of the
322: synchronizable ability.
323:
324: Secondly, we implement the same measures on the SF network. The
325: model \cite{SF} is generated by randomly connecting nodes, forcing
326: each nodes $i$ to have connectivity $k_i\geq k_0$ which follows the
327: probability distribution $P(k)\sim k^{-\gamma}$. Note that
328: decreasing $\gamma$ increases $S$ by inducing the longer tail in the
329: connectivity distribution and smaller $\gamma$ gives rise to shorter
330: $D$, as exhibited in the upper layer of Fig. \ref{fig:swsf}(b). And
331: in its lower layer, $R$ is observed to increase as $\gamma$ is
332: increased, which indicates less synchronized when less
333: heterogeneous.
334: % and differs from the results in (T. Nishikawa, A.E.Motter, Y-C Lai, and F.C. Hoppensteadt PRL \textbf{91}, 014101 (2003)).
335:
336: The results shown in Fig. \ref{fig:swsf} imply that the
337: synchronizable ability on the SW and SF networks is improved as the
338: heterogeneity of the degree distribution is increased or as the
339: average network distance is decreased. In other words, as the
340: network becomes more heterogeneous and ``smaller", with $S$ gets
341: larger and $D$ gets shorter, it would becomes synchronized more
342: closely by the collective dynamics.
343:
344: In summary, we have investigated the synchronization dynamics of the
345: KM departing from the fully synchronized state. We have got the
346: analytic expression of the dynamical correlation between pairs of
347: oscillators that reveals the relation between the network dynamics
348: and the underlying topology. Moreover, it also reveal the internal
349: structure of networks that can be used as a new algorithm to detect
350: community structures. Further, we have proposed a new measure about
351: the synchronization in complex networks and scrutinize it in SW and
352: SF networks. Our results indicate that the more heterogeneous and
353: ``smaller" the network is, the more closely it would be synchronized
354: by the collective dynamics. We hope our study can provide new
355: insights into the understanding of the role of synchronization
356: between the network structure and function. We also expect it can
357: provide new tools to detect community structure and to analyze the
358: ubiquitous synchronization phenomenon.
359:
360:
361:
362: \begin{thebibliography}{10}
363: \bibitem{review1} S. Strogatz, Nature (London) \textbf{410}, 268 (2001).
364: \bibitem{review2} R. Albert and A.-L. Barabasi, Rev. Mod. Phys. \textbf{74}, 47 (2002).
365: \bibitem{review3} M. E. J. Newman, SIAM Rev. \textbf{45}, 167 (2003).
366: \bibitem{review4} S. Boccaletti, V. Latora, Y. Moreno, M. Chavez,
367: and D.-U. Hwang, Phys. Rep. \textbf{424}, 175 (2006).
368:
369: \bibitem{sync1} S. H. Strogatz, {\sl Sync: The Emerging Science of
370: Spontaneous Order} (Hyperion, New York, 2003).
371: \bibitem{sync2} M. Barahona and L. M. Pecora, Phys. Rev. Lett. \textbf{89}, 054101 (2002).
372: \bibitem{sync3} T. Nishikawa, A. E. Motter, Y.-C. Lai, and F. C. Hoppensteadt, Phys. Rev. Lett. \textbf{91}, 014101 (2003).
373: \bibitem{sync4} H. Hong, B. J. Kim, M. Y. Choi, and H. Park, Phys. Rev. E \textbf{69}, 067105 (2004).
374: \bibitem{sync5} Y. Moreno and A. F. Pacheco, Europhys. Lett. \textbf{68}, 603 (2004).
375: \bibitem{sync6} A. E. Motter, C. Zhou, and J. Kurths, Phys. Rev. E \textbf{71}, 016116 (2005).
376: \bibitem{sync7} E. Oh, K. Rho, H. Hong, and B. Kahng, Phys. Rev. E \textbf{72}, 047101 (2005).
377: \bibitem{sync8} L. Donetti, P. I. Hurtado, and M. A. Mu\~{n}oz, Phys. Rev. Lett. \textbf{95}, 188701 (2005).
378: \bibitem{sync9} M. Chavez, D.-U. Hwang, A. Amann, H. G. E. Hentschel, and S. Boccaletti, Phys. Rev. Lett. \textbf{94}, 218701 (2005).
379: \bibitem{sync10} J. G. Restrepo, E. Ott, and B. R. Hunt, Chaos \textbf{16}, 015107 (2006).
380: \bibitem{sync11} C. Zhou and J. Kurths, Chaos \textbf{16},015104 (2006)
381: \bibitem{sync12} A. Arenas, A. Diaz-Guilera, and C. J. Perez-Vicente, Phys. Rev. Lett. \textbf{96}, 114102 (2006).
382: \bibitem{sync13} J. G\'{o}mez-Garde\~{n}es, Y. Moreno, and Alex Arenas, Phys. Rev. Lett. \textbf{98}, 034101 (2007).
383:
384: \bibitem{KM} Y. Kuramoto, {\sl Chemical Oscillations, Waves and Turbulence}
385: (Springer-Verlag, Berlin, 1984).
386:
387: \bibitem{KM1} J. A. Acebron, L. L. Bonilla, C. J. Perez Vicente, F.
388: Ritort, and R. Spigler, Rev. Mod. Phys. \textbf{77}, 137 (2005).
389:
390: \bibitem{MSF} L. M. Pecora and T. L. Carroll, Phys. Rev. Lett. \textbf{80}, 2109 (1998).
391:
392: \bibitem{method} N. G. Van Kampen, {\sl Stochastic Processes in Physics and
393: Chemistry} (North-Holland, Amsterdam, 1981).
394:
395: \bibitem{Girvan} M. Girvan and M. E. J. Newman PNAS \textbf{99}, 7821 (2002).
396:
397: \bibitem{WS} D. J. Watts and S. H. Strogatz, Nature (London) \textbf{393}, 440 (1998).
398:
399: \bibitem{SF} M. E. J. Newman, S. H. Strogatz, and D. J. Watts, Phys. Rev. E \textbf{64}, 026118 (2001).
400:
401: \end{thebibliography}
402: \end{document}
403: