Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex...

18
Exploring Synchronization in Complex Oscillator Networks Florian D¨ orfler Francesco Bullo Abstract— The emergence of synchronization in a network of coupled oscillators is a pervasive topic in various scientific disciplines ranging from biology, physics, and chemistry to social networks and engineering applications. A coupled oscilla- tor network is characterized by a population of heterogeneous oscillators and a graph describing the interaction among the oscillators. These two ingredients give rise to a rich dynamic behavior that keeps on fascinating the scientific community. In this article, we present a tutorial introduction to coupled oscillator networks, we review the vast literature on theory and applications, and we present a collection of different synchro- nization notions, conditions, and analysis approaches. We focus on the canonical phase oscillator models occurring in countless real-world synchronization phenomena, and present their rich phenomenology. We review a set of applications relevant to control scientists. We explore different approaches to phase and frequency synchronization, and we present a collection of synchronization conditions and performance estimates. For all results we present self-contained proofs that illustrate a sample of different analysis methods in a tutorial style. I. I NTRODUCTION The scientific interest in synchronization of coupled oscil- lators can be traced back to the work by Christiaan Huygens on “an odd kind sympathy” between coupled pendulum clocks [1], and it still fascinates the scientific community nowadays [2], [3]. Within the rich modeling phenomenology on synchronization among coupled oscillators, we focus on the canonical model of a continuous-time limit-cycle oscil- lator network with continuous and bidirectional coupling. A network of coupled phase oscillators: A mechanical analog of a coupled oscillator network is the spring network shown in Figure 1 and consists of a group of kinematic particles constrained to rotate around a circle and assumed to move without colliding. Each particle is characterized by ω 1 ω 3 ω 2 a 12 a 13 a 23 Fig. 1. Mechanical analog of a coupled oscillator network This material is based in part upon work supported by NSF grants IIS- 0904501 and CPS-1135819. Florian D¨ orfler and Francesco Bullo are with the Center for Control, Dynamical Systems and Computation, University of California at Santa Barbara. Email: {dorfler,bullo}@engineering.ucsb.edu a phase angle θ i S 1 and has a preferred natural rotation frequency ω i R. Pairs of interacting particles i and j are coupled through an elastic spring with stiffness a ij > 0. We refer to the Appendix A for a first principle modeling of the spring-interconnected particles depicted in Figure 1. Formally, each isolated particle is an oscillator with first- order dynamics ˙ θ i = ω i . The interaction among n such oscillators is modeled by a connected graph G(V , E ,A) with nodes V = {1,...,n}, edges E ⊂V×V , and positive weights a ij > 0 for each undirected edge {i, j }∈E . Under these assumptions, the overall dynamics of the coupled oscillator network are ˙ θ i = ω i - X n j=1 a ij sin(θ i - θ j ) , i ∈{1,...,n} . (1) The rich dynamic behavior of the coupled oscillator model (1) arises from a competition between each oscillator’s tendency to align with its natural frequency ω i and the synchronization-enforcing coupling a ij sin(θ i - θ j ) with its neighbors. Intuitively, a weakly coupled and strongly het- erogeneous network does not display any coherent behavior, whereas a strongly coupled and sufficiently homogeneous network is amenable to synchronization, where all frequen- cies ˙ θ i (t) or even all phases θ i (t) become aligned. History, applications and related literature: The cou- pled oscillator model (1) has first been proposed by Arthur Winfree [4]. In the case of a complete interaction graph, the coupled oscillator dynamics (1) are nowadays known as the Kuramoto model of coupled oscillators due to Yoshiki Kuramoto [5], [6]. Stephen Strogatz provides an excellent historical account in [7]. We also recommend the survey [8]. Despite its apparent simplicity, the coupled oscillator model (1) gives rise to rich dynamic behavior. This model is encountered in various scientific disciplines ranging from natural sciences over engineering applications to social net- works. The model and its variations appear in the study if biological synchronization phenomena such as pacemaker cells in the heart [9], circadian rhythms [10], neuroscience [11]–[13], metabolic synchrony in yeast cell populations [14], flashing fireflies [15], chirping crickets [16], biological locomotion [17], animal flocking behavior [18], fish schools [19], and rhythmic applause [20], among others. The coupled oscillator model (1) also appears in physics and chemistry in modeling and analysis of spin glass models [21], [22], flavor evolutions of neutrinos [23], coupled Josephson junctions [24], and in the analysis of chemical oscillations [25]. Some technological applications of the coupled oscillator model (1) include deep brain stimulation [26], [27], vehicle coordination [19], [28]–[31], carrier synchronization without arXiv:1209.1335v1 [math.OC] 6 Sep 2012

Transcript of Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex...

Page 1: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Exploring Synchronization in Complex Oscillator Networks

Florian Dorfler Francesco Bullo

Abstract— The emergence of synchronization in a networkof coupled oscillators is a pervasive topic in various scientificdisciplines ranging from biology, physics, and chemistry tosocial networks and engineering applications. A coupled oscilla-tor network is characterized by a population of heterogeneousoscillators and a graph describing the interaction among theoscillators. These two ingredients give rise to a rich dynamicbehavior that keeps on fascinating the scientific community.In this article, we present a tutorial introduction to coupledoscillator networks, we review the vast literature on theory andapplications, and we present a collection of different synchro-nization notions, conditions, and analysis approaches. We focuson the canonical phase oscillator models occurring in countlessreal-world synchronization phenomena, and present their richphenomenology. We review a set of applications relevant tocontrol scientists. We explore different approaches to phaseand frequency synchronization, and we present a collection ofsynchronization conditions and performance estimates. For allresults we present self-contained proofs that illustrate a sampleof different analysis methods in a tutorial style.

I. INTRODUCTION

The scientific interest in synchronization of coupled oscil-lators can be traced back to the work by Christiaan Huygenson “an odd kind sympathy” between coupled pendulumclocks [1], and it still fascinates the scientific communitynowadays [2], [3]. Within the rich modeling phenomenologyon synchronization among coupled oscillators, we focus onthe canonical model of a continuous-time limit-cycle oscil-lator network with continuous and bidirectional coupling.

A network of coupled phase oscillators: A mechanicalanalog of a coupled oscillator network is the spring networkshown in Figure 1 and consists of a group of kinematicparticles constrained to rotate around a circle and assumedto move without colliding. Each particle is characterized by

x

x

x

ω1

ω3ω2

a12

a13

a23

Fig. 1. Mechanical analog of a coupled oscillator network

This material is based in part upon work supported by NSF grants IIS-0904501 and CPS-1135819.

Florian Dorfler and Francesco Bullo are with the Center for Control,Dynamical Systems and Computation, University of California at SantaBarbara. Email: dorfler,[email protected]

a phase angle θi ∈ S1 and has a preferred natural rotationfrequency ωi ∈ R. Pairs of interacting particles i and j arecoupled through an elastic spring with stiffness aij > 0. Werefer to the Appendix A for a first principle modeling of thespring-interconnected particles depicted in Figure 1.

Formally, each isolated particle is an oscillator with first-order dynamics θi = ωi. The interaction among n suchoscillators is modeled by a connected graph G(V, E , A) withnodes V = 1, . . . , n, edges E ⊂ V × V , and positiveweights aij > 0 for each undirected edge i, j ∈ E . Underthese assumptions, the overall dynamics of the coupledoscillator network are

θi = ωi−∑n

j=1aij sin(θi− θj) , i ∈ 1, . . . , n . (1)

The rich dynamic behavior of the coupled oscillator model(1) arises from a competition between each oscillator’stendency to align with its natural frequency ωi and thesynchronization-enforcing coupling aij sin(θi − θj) with itsneighbors. Intuitively, a weakly coupled and strongly het-erogeneous network does not display any coherent behavior,whereas a strongly coupled and sufficiently homogeneousnetwork is amenable to synchronization, where all frequen-cies θi(t) or even all phases θi(t) become aligned.

History, applications and related literature: The cou-pled oscillator model (1) has first been proposed by ArthurWinfree [4]. In the case of a complete interaction graph,the coupled oscillator dynamics (1) are nowadays known asthe Kuramoto model of coupled oscillators due to YoshikiKuramoto [5], [6]. Stephen Strogatz provides an excellenthistorical account in [7]. We also recommend the survey [8].

Despite its apparent simplicity, the coupled oscillatormodel (1) gives rise to rich dynamic behavior. This modelis encountered in various scientific disciplines ranging fromnatural sciences over engineering applications to social net-works. The model and its variations appear in the study ifbiological synchronization phenomena such as pacemakercells in the heart [9], circadian rhythms [10], neuroscience[11]–[13], metabolic synchrony in yeast cell populations[14], flashing fireflies [15], chirping crickets [16], biologicallocomotion [17], animal flocking behavior [18], fish schools[19], and rhythmic applause [20], among others. The coupledoscillator model (1) also appears in physics and chemistry inmodeling and analysis of spin glass models [21], [22], flavorevolutions of neutrinos [23], coupled Josephson junctions[24], and in the analysis of chemical oscillations [25].

Some technological applications of the coupled oscillatormodel (1) include deep brain stimulation [26], [27], vehiclecoordination [19], [28]–[31], carrier synchronization without

arX

iv:1

209.

1335

v1 [

mat

h.O

C]

6 S

ep 2

012

Page 2: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

phase-locked loops [32], semiconductor lasers [33], [34], mi-crowave oscillators [35], clock synchronization in decentral-ized computing networks [36]–[41], decentralized maximumlikelihood estimation [42], and droop-controlled inverters inmicrogrids [43]. Finally, the coupled oscillator model (1)also serves as the prototypical example for synchronizationin complex networks [44]–[47] and its linearization is thewell-known consensus protocol studied in networked control,see the surveys and monographs [48]–[50]. Various controlscientists explored the coupled oscillator model (1) as anonlinear generalization of the consensus protocol [51]–[57].

Second-order variations of the coupled oscillator model(1) appear in synchronization phenomena, in population offlashing fireflies [58], in particle models mimicking animalflocking behavior [59], [60], in structure-preserving powersystem models, [61], [62] in network-reduced power systemmodels [63], [64], in coupled metronomes [65], in pedes-trian crowd synchrony on London’s Millennium bridge [66],and in Huygen’s pendulum coupled clocks [67]. Coupledoscillator networks with second-order dynamics have beentheoretically analyzed in [8], [68]–[74], among others.

Coupled oscillator models of the form (1) are also studiedfrom a purely theoretic perspective in the physics, dynamicalsystems, and control communities. At the heart of the cou-pled oscillator dynamics is the transition from incoherenceto synchrony. Here, different notions and degrees of synchro-nization can be distinguished [74]–[76], and the (apparently)incoherent state features rich and largely unexplored dynam-ics as well [47], [77]–[79]. In this article we will be particu-larly interested in phase and frequency synchronization whenall phases θi(t) become aligned, respectively all frequenciesθi(t) become aligned. We refer to [7], [8], [19], [28], [31],[52], [53], [56], [64], [74]–[76], [80]–[95], [95]–[114] for anincomplete overview concerning numerous recent researchactivities. We will review some of literature throughout thepaper and refer to the surveys [7], [8], [44]–[46], [74] forfurther applications and numerous additional theoretic resultsconcerning the coupled oscillator model (1).

Contributions and contents: In this paper, we introducethe reader to synchronization in networks of coupled oscil-lators. We present a sample of important analysis conceptsin a tutorial style and from a control-theoretic perspective.

In Section II, we will review a set of selected technologicalapplications which are directly tied to the coupled oscillatormodel (1) and also relevant to control systems. We will covervehicle coordination, electric power networks, and clocksynchronization in depth, and also justify the importanceof the coupled oscillator model (1) as a canonical model.Prompted by these applications, we review the existing re-sults concerning phase synchronization, phase balancing, andfrequency synchronization, and we also present some novelresults on synchronization in sparsely-coupled networks.

In particular, Section III introduces the reader to dif-ferent synchronization notions, performance metrics, andsynchronization conditions. We illustrate these results witha simple yet rich example that nicely explains the basicphenomenology in coupled oscillator networks.

Section IV presents a collection of important resultsregarding phase synchronization, phase balancing, and fre-quency synchronization. By now the analysis methods forsynchronization have reached a mature level, and we presentsimple and self-contained proofs using a sample of differentanalysis methods. In particular, we present one result onphase synchronization and one result on phase balancing in-cluding estimates on the exponential synchronization rate andthe region of attraction (see Theorem 4.3 and Theorem 4.4).We also present some implicit and explicit, and necessaryand sufficient conditions for frequency synchronization inthe classic homogeneous case of a complete and uniformly-weighted coupling graphs (see Theorem 4.5). Concerningfrequency synchronization in sparse graphs, we present twopartially new synchronization conditions depending on thealgebraic connectivity (see Theorem 4.6 and Theorem 4.7).

In our technical presentation, we try to strike a balancebetween mathematical precision and removing unnecessarytechnicalities. For this reason some proofs are reported in theappendix and others are only sketched here with referencesto the detailed proofs elsewhere. Hence, the main technicalideas are conveyed while the tutorial value is maintained.

Finally, Section V concludes the paper. We summarize thelimitations of existing analysis methods and suggest someimportant directions for future research.

Preliminaries and notation: The remainder of this sec-tion introduces some notation and recalls some preliminaries.

Vectors and functions: Let 1n and 0n be the n-dimensionalvector of unit and zero entries, and let 1⊥n be the orthogonalcomplement of 1n in Rn, that is, 1⊥n , x ∈ Rn : x ⊥ 1n.Given an n-tuple (x1, . . . , xn), let x ∈ Rn be the associatedvector with maximum and minimum elements xmax and xmin.For an ordered index set I of cardinality |I| and an one-dimensional array xii∈I , let diag(cii∈I) ∈ R|I|×|I| bethe associated diagonal matrix. Finally, define the continuousfunction sinc : R→ R by sinc(x) = sin(x)/x for x 6= 0.

Geometry on the n-torus: The set S1 denotes the unitcircle, an angle is a point θ ∈ S1, and an arc is a connectedsubset of S1. The geodesic distance between two anglesθ1, θ2 ∈ S1 is the minimum of the counter-clockwise andthe clockwise arc lengths connecting θ1 and θ2. With slightabuse of notation, let |θ1 − θ2| denote the geodesic distancebetween two angles θ1, θ2 ∈ S1. The n-torus is the productset Tn = S1 × · · · × S1 is the direct sum of n unit circles.For γ ∈ [0, 2π[, let Arcn(γ) ⊂ Tn be the closed set of anglearrays θ = (θ1, . . . , θn) with the property that there existsan arc of length γ containing all θ1, . . . , θn. Thus, an anglearray θ ∈ Arcn(γ) satisfies maxi,j∈1,...,n |θi − θj | ≤ γ.Finally, let Arcn(γ) be the interior of the set Arcn(γ).

Algebraic graph theory: Let G(V, E , A) be an undirected,connected, and weighted graph without self-loops. Let A ∈Rn×n be its symmetric nonnegative adjacency matrix withzero diagonal, aii = 0. For each node i ∈ 1, . . . , n, definethe nodal degree by degi =

∑nj=1 aij . Let L ∈ Rn×n be

the Laplacian matrix defined by L = diag(degini=1)−A.If a number ` ∈ 1, . . . , |E| and an arbitrary direction isassigned to each edge i, j ∈ E , the (oriented) incidence

Page 3: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

matrix B ∈ Rn×|E| is defined component-wise by Bk` = 1if node k is the sink node of edge ` and by Bk` = −1 ifnode k is the source node of edge `; all other elements arezero. For x ∈ Rn, the vector BTx has components xi − xjcorresponding to the oriented edge from j to i, that is, BT

maps node variables xi, xj to incremental edge variablesxi−xj . If diag(aiji,j∈E) is the diagonal matrix of edgeweights, then L = B diag(aiji,j∈E)BT . If the graphis connected, then Ker (BT ) = Ker (L) = span(1n), alln − 1 non-zero eigenvalues of L are strictly positive, andthe second-smallest eigenvalue λ2(L) is called the algebraicconnectivity and is a spectral connectivity measure.

II. APPLICATIONS OF KURAMOTO OSCILLATORSRELEVANT TO CONTROL SYSTEMS

The mechanical analog in Figure 1 provides an intuitiveillustration of the coupled oscillator dynamics (1), and wereviewed a wide range of examples from physics, life sci-ences, and technology in Section I. Here, we detail a setof selected technological applications which are relevant tocontrol systems scientists.

A. Flocking, Schooling, and Planar Vehicle Coordination

An emerging research field in control is the coordination ofautonomous vehicles based on locally available informationand inspired by biological flocking phenomena. Consider aset of n particles in the plane R2, which we identify withthe complex plane C. Each particle i ∈ V = 1, . . . , nis characterized by its position ri ∈ C, its heading angleθi ∈ S1, and a steering control law ui(r, θ) depending onthe position and heading of itself and other vehicles. Forsimplicity, we assume that all particles have constant andunit speed. The particle kinematics are then given by [115]

ri = eiθi ,

θi = ui(r, θ) ,

i ∈ 1, . . . , n , (2)

where i =√−1 is the imaginary unit. If the control ui is

identically zero, then particle i travels in a straight line withorientation θi(0), and if ui = ωi ∈ R is a nonzero constant,then the particle traverses a circle with radius 1/|ωi|.

The interaction among the particles is modeled by apossibly time-varying interaction graph G(V, E(t), A(t)) de-termined by communication and sensing patterns. Someinteresting motion patterns emerge if the controllers use onlyrelative phase information between neighboring particles,that is, ui = ω0(t) + fi(θi − θj) for i, j ∈ E(t) andω0 : R≥0 → R. For example, the control ui = ω0(t) −K ·∑n

j=1 aij(t) sin(θi − θj) with gain K ∈ R results in

θi = ω0(t)−K ·∑n

j=1aij(t) sin(θi − θj) , i ∈ V . (3)

The controlled phase dynamics (3) correspond to the coupledoscillator model (1) with a time-varying interaction graphwith weights K · aij(t) and identically time-varying naturalfrequencies ωi = ω0(t) for all i ∈ 1, . . . , n. The controlledphase dynamics (3) give rise to very interesting coordinationpatterns that mimic animal flocking behavior [18] and fish

schools [19]. Inspired by these biological phenomena, thecontrolled phase dynamics (3) and its variations have alsobeen studied in the context of tracking and formation con-trollers in swarms of autonomous vehicles [19], [28]–[31].A few trajectories are illustrated in Figure 2, and we refer to[19], [28]–[31] for other control laws and motion patterns.

In the following sections, we will present various tools toanalyze the motion patterns in Figure 2, which we will referto as phase synchronization and phase balancing.

(a) (b)

Fig. 2. Illustration of the controlled dynamics (2)-(3) with n=6 particles,a complete interaction graph, and identical and constant natural frequenciesω0(t) = 1, where K = 1 in panel (a) and K = −1 in panel (b). Thearrows depict the orientation, the dashed curves show the long-term positiondynamics, and the solid curves show the initial transient position dynamics.It can be seen that even for this simple choice of controller, the resultingmotion results in “synchronized” or “balanced” heading angles forK = ±1.

B. Power Grids with Synchronous Generators and Inverters

Here, we present the structure-preserving power networkmodel introduced in [61] and refer to [62, Chapter 7] fordetailed derivation from a higher order first principle model.Additionally, we equip the power network model with a setof inverters and refer to [43] for a detailed modeling.

Consider an alternating current (AC) power network mod-eled as an undirected, connected, and weighted graph withnode set V = 1, . . . , n, transmission lines E ⊂ V ×V , andadmittance matrix Y =Y T ∈ Cn×n. For each node, considerthe voltage phasor Vi = |Vi|eiθi corresponding to the phaseθi ∈ S1 and magnitude |Vi| ≥ 0 of the sinusoidal solutionto the circuit equations. If the network is lossless, then theactive power flow from node i to j is aij sin(θi−θj), wherewe used the shorthand aij = |Vi| · |Vj | · =(Yij).

In the following, we assume that the node set is partitionedas V = V1 ∪V2 ∪V3, where V1 are load buses, V2 are con-ventional synchronous generators, and V3 are grid-connecteddirect current (DC) power sources, such as solar farms. Theactive power drawn by a load i ∈ V1 consists of a constantterm Pl,i > 0 and a frequency-dependent term Diθi withDi > 0. The resulting power balance equation is

Diθi + Pl,i = −∑n

j=1aij sin(θi − θj) , i ∈ V1 . (4)

If the generator reactances are absorbed into the admittancematrix, then the swing dynamics of generator i ∈ V2 are

Miθi+Diθi = Pm,i−∑n

j=1aij sin(θi−θj) , i ∈ V2, (5)

Page 4: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

where θi ∈ S1 and θi ∈ R1 are the generator rotor angleand frequency, Pm,i > 0 is the mechanical power input, andMi > 0, and Di > 0 are the inertia and damping coefficients.

We assume that each DC source is connected to the ACgrid via an DC/AC inverter, the inverter output impendancesare absorbed into the admittance matrix, and each inverter isequipped with a conventional droop-controller. For a droop-controlled inverter i ∈ V3 with droop-slope 1/Di > 0, thedeviation of the power output

∑nj=1 aij sin(θi − θj) from

its nominal value Pd,i > 0 is proportional to the frequencydeviation Diθi. This gives rise to the inverter dynamics

Diθi = Pd,i −∑n

j=1aij sin(θi − θj) , i ∈ V3 . (6)

These power network devices are illustrated in Figure 3.Finally, we remark that different load models such as con-

Pm,i |Vi| eii Yij

|Vi| eii

YijYik

DiPl,i

(a) (b)

(c)

(d)

|Vj | eijYij|Vi| eii

aij sin(i j)

+

aij sin(i j)

Pd,i

|Vi| eii

Fig. 3. Illustration of the power network devices as circuit elements. Sub-figure (a) shows a transmission element connecting nodes i and j, Subfigure(b) shows a frequency-dependent load, Subfigure (c) shows an inverter con-trolled according to (6), and Subfigure (d) shows a synchronous generator.

stant power/current/susceptance loads and synchronous mo-tor loads can be modeled and analyzed by the same set ofequations (4)-(6), see [62]–[64], [116], [117].

Synchronization is pervasive in the operation of powernetworks. All generating units of an interconnected grid mustremain in strict frequency synchronism while continuouslyfollowing demand and rejecting disturbances. Notice that,with exception of the inertial terms Miθi and the possiblynon-unit coefficients Di, the power network dynamics (4)-(6)are a perfect electrical analog of the coupled oscillator model(1) with ω = (−Pl,i, Pm,i, Pd,i). Thus, it is not surprisingthat scientists from different disciplines recently advocatedcoupled oscillator approaches to analyze synchronization inpower networks [43], [64], [69], [97], [114], [118]–[122].

The theoretic tools presented in the following sectionsestablish how frequency synchronization in power networksdepend on the nodal parameters (Pl,i, Pm,i, Pd,i) as wellas the interconnecting electrical network with weights aij .Ultimately, this deep understanding of synchrony gives usthe correct intuition to design controllers and remedial actionschemes preventing the loss of synchrony.

C. Clock Synchronization in Decentralized Networks

Another emerging technological application of the coupledoscillator model (1) is clock synchronization in decentralizedcomputing networks, such as wireless and distributed soft-ware networks. A natural approach to clock synchronizationis to treat each clock as a coupled oscillator and follow adiffusion-based protocol to synchronize them, see the historicand recent surveys [36], [37], the landmark paper [38], andthe interesting recent results [39]–[41].

Consider a set of distributed processors V = 1, . . . , ninterconnected in a (possibly directed) communication net-work. Each processor is equipped with an internal softwareclock, and these clocks need to be synchronized for dis-tributed computing and network routing tasks. For simplicity,we consider only analog clocks with continuous couplingsince digital clocks are essentially discretized analog clocksand pulse-coupled clocks can be modeled continuously aftera phase reduction and averaging analysis.

For our purposes, the clock of processor i is a voltage-controlled oscillator which outputs a harmonic waveformsi(t) = sin(θi(t)), where θi(t) is the accumulated instanta-neous phase. For uncoupled nodes, the phase θi(t) evolves as

θi(t) =

(θi(0) +

Tnom + Tit

)mod(2π) , i ∈ 1, . . . , n .

where Tnom > 0 is the nominal period, Ti ∈ R is an offset(frequency offset or skew), and θi(0) ∈ S1 is the initialphase. To synchronize their internal clocks, the processorsfollow a diffusion-based protocol. In a first step, neighboringoscillators continuously communicate their respective wave-forms si(t) to another. Second, through a phase detector eachnode measures a convex combination of phase differences as

cvxi(θ(t)) =∑n

j=1aijf(θi(t)− θj(t)) , i ∈ 1, . . . , n ,

where aij ≥ 0 are convex (∑nj=1 aij = 1) and detector-

specific weights, and f : S1 → R is an odd 2π-periodic func-tion. Finally, cvxi(θ(t)) is fed to a (first-order and constant)phase-locked loop filter K whose output drives the localphase according to

θi(t) =2π

Ti+K · cvxi(θ(t)) , i ∈ 1, . . . , n . (7)

The goal of the synchronization protocol (7) is to synchronizethe frequencies θi(t) or even the phases θi(t) in the processornetwork. For an undirected communication protocol, sym-metric weights aij = aji, and a sinusoidal coupling functionf(·) = sin(·), the synchronization protocol (7) equals againthe coupled oscillator model (1).

The tools developed in the next section will enable us tostate conditions when the protocol (7) successfully achievesphase or frequency synchronization. Of course, the protocol(7) is merely a starting point, more sophisticated phase-locked loop filters can be constructed to enhance steady-statedeviations from synchrony, and communication and phasenoise as well as time-delays can be considered in the design.

Page 5: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

D. Canonical Coupled Oscillator Model

The importance of the coupled oscillator model (1) doesnot stem only from the various examples listed in Sections Iand II. Even though model (1) appears to be quite specific (aphase oscillator with constant driving term and continuous,diffusive, and sinusoidal coupling), it is the canonical modelof coupled limit-cycle oscillators [123]. In the following,we briefly sketch how such general models can be re-duced to model (1). We schematically follow the approaches[124, Chapter 10], [125] developed in the computationalneuroscience community without aiming at mathematicalprecision, and we refer to [123], [126] for further details.

Consider an oscillator modeled as a dynamical system withstate x ∈ Rm and nonlinear dynamics x = f(x), whichadmit a locally exponentially stable periodic orbit γ ⊂ Rmwith period T > 0. By a change of variables, any trajectoryin a local neighborhood of γ can be characterized by a phasevariable ϕ ∈ S1 with dynamics ϕ = Ω, where Ω = 2π/T .

Now consider a weakly forced oscillator of the form

x = f(x) + ε · δ(t) , (8)

where ε > 0 is sufficiently small and δ(t) is a time-dependentforcing term. For small forcing εδ(t), the attractive limitcycle γ persists, and the phase dynamics are obtained as

ϕ = Ω + εQ(ϕ)δ(t) +O(ε2) ,

where Q(ϕ) is the infinitesimal phase response curve (orlinear response function), and we dropped higher order terms.

Now consider n such limit cycle oscillators, where xi ∈Rm is the state of oscillator i with limit cycle γi ⊂ Rm andperiod Ti > 0. We assume that the oscillators are weaklycoupled with interaction graph G(V, E) and dynamics

xi = fi(xi)+ε∑i,j∈E

gij(xi, xj) , i ∈ 1, . . . , n , (9)

where gij(·) is the coupling function for the pair i, j ∈ E .The coupling gij(·) can possibly be impulsive. The weakcoupling in (9) can be identified with the weak forcing in(8), and a transformation to phase coordinates yields

ϕi = Ωi + ε∑i,j∈E

Qi(ϕ)gij(xi(ϕi), xj(ϕj)) ,

where Ωi = 2π/Ti. The local change of variables θi(t) =ϕi(t)− Ωit then yields the coupled phase dynamics

θi = ε∑i,j∈E

Qi(θi+Ωit)gij(xi(θi+Ωit), xj(θj+Ωjt)).

An averaging analysis applied to the θ-dynamics results in

θi = εωi + ε∑i,j∈E

hij(θi − θj) , (10)

where ωi = hii(0) and the averaged coupling functions are

hij(χ) = limT→∞

1

T

∫ T

0

Qi(Ωiτ)gij(xi(Ωiτ), xj(Ωjτ−χ))dτ.

Notice that the averaged coupling functions hij are 2π-periodic and the coupling is diffusive. If all functions hijare odd, a first-order Fourier series expansion of hij yieldshij(·) ≈ aij sin(·) as first harmonic with some coefficient

aij . In this case, the dynamics (10) in the slow time scaleτ = εt reduce exactly to the coupled oscillator model (1).

This analysis justifies calling the coupled oscillator model(1) the canonical model for coupled limit-cycle oscillators.

III. SYNCHRONIZATION NOTIONS AND METRICS

In this section, we introduce different notions of syn-chronization. Whereas the first four subsections address thecommonly studied notions of synchronization associatedwith a coherent behavior and cohesive phases, SubsectionIII-E addresses the converse concept of phase balancing.

A. Synchronization Notions

The coupled oscillator model (1) evolves on Tn, andfeatures an important symmetry, namely the rotational in-variance of the angular variable θ. This symmetry gives riseto the rich synchronization dynamics. Different levels of syn-chronization can be distinguished, and the most commonlystudied notions are phase and frequency synchronization.

Phase synchronization: A solution θ : R≥0 → Tn to thecoupled oscillator model (1) achieves phase synchronizationif all phases θi(t) become identical as t→∞.

Phase cohesiveness: As we will see later, phase synchro-nization can occur only if all natural frequencies ωi areidentical. If the natural frequencies are not identical, theneach pairwise distance |θi(t) − θj(t)| can converge to aconstant but not necessarily zero value. The concept of phasecohesiveness formalizes this possibility. For γ ∈ [0, π[, let∆G(γ) ⊂ Tn be the closed set of angle arrays (θ1, . . . , θn)with the property |θi−θj | ≤ γ for all i, j ∈ E , that is, eachpairwise phase distance is bounded by γ. Also, let ∆G(γ) bethe interior of ∆G(γ). Notice that Arcn(γ) ⊆ ∆G(γ) but thetwo sets are generally not equal. A solution θ : R≥0 → Tnis then said to be phase cohesive if there exists a lengthγ ∈ [0, π[ such that θ(t) ∈ ∆G(γ) for all t ≥ 0.

Frequency synchronization: A solution θ : R≥0 → Tnachieves frequency synchronization if all frequencies θi(t)converge to a common frequency ωsync ∈ R as t→∞. Theexplicit synchronization frequency ωsync ∈ R of the coupledoscillator model (1) can be obtained by summing over allequations in (1) as

∑ni=1 θi =

∑ni=1 ωi. In the frequency-

synchronized case, this sum simplifies to∑ni=1 ωsync =∑n

i=1 ωi. In conclusion, if a solution of the coupled oscillatormodel (1) achieves frequency synchronization, then it does sowith synchronization frequency equal to ωsync =

∑ni=1 ωi/n.

By transforming to a rotating frame with frequency ωsync andby replacing ωi by ωi−ωsync, we obtain ωsync = 0 (or equiv-alently ω ∈ 1⊥n ). In what follows, without loss of generality,we will sometimes assume that ω ∈ 1⊥n so that ωsync = 0.

Remark 1 (Terminology): Alternative terminologies forphase synchronization include full, exact, or perfect synchro-nization. For a frequency-synchronized solution all phasedistances |θi(t)− θj(t)| are constant in a rotating coordinateframe with frequency ωsync, and the terminology phaselocking is sometimes used instead of frequency synchroniza-tion. Other commonly used terms include frequency locking,frequency entrainment, or also partial synchronization.

Page 6: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Synchronization: The main object under study in mostapplications and theoretic analyses are phase cohesive andfrequency-synchronized solutions, that is, all oscillators ro-tate with the same synchronization frequency, and all theirpairwise phase distances are bounded. In the following, werestrict our attention to synchronized solutions with suffi-ciently small phase distances |θi−θj | ≤ γ < π/2 for i, j ∈E . Of course, there may exist other possible solutions, butthese are not necessarily stable (see our analysis in SectionIV) or not relevant in most applications1. We say that asolution θ : R≥0 → Tn to the coupled oscillator model(1) is synchronized if there exists θsync ∈ ∆G(γ) for someγ ∈ [0, π/2[ and ωsync ∈ R (identically zero for ω ∈ 1⊥n )such that θ(t) = θsync + ωsync1nt (mod 2π) for all t ≥ 0.

Synchronization manifold: The geometric object understudy in synchronization is the synchronization manifold.Given a point r ∈ S1 and an angle s ∈ [0, 2π], let rots(r) ∈S1 be the rotation of r counterclockwise by the angle s. For(r1, . . . , rn) ∈ Tn, define the equivalence class

[(r1, . . . , rn)]=(rots(r1), . . . , rots(rn)) ∈ Tn |s ∈ [0, 2π].Clearly, if (r1, . . . , rn) ∈ ∆G(γ) for some γ ∈ [0, π/2[,then [(r1, . . . , rn)] ⊂ ∆G(γ). Given a synchronized solutioncharacterized by θsync ∈ ∆G(γ) for some γ ∈ [0, π/2[,the set [θsync] ⊂ ∆G(γ) is a synchronization manifold ofthe coupled-oscillator model (1). Note that a synchronizedsolution takes value in a synchronization manifold due torotational symmetry, and for ω ∈ 1⊥n (implying ωsync = 0) asynchronization manifold is also an equilibrium manifold ofthe coupled oscillator model (1). These geometric conceptsare illustrated in Figure 4 for the two-dimensional case.

∆G(π/2)

[θ∗]

12

θ∗

Fig. 4. Illustration of the state space T2, the set ∆G(π/2), the synchro-nization manifold [θ∗] associated to a phase-synchronized angle array θ∗ =(θ∗1 , θ

∗2) ∈ ∆G(0), and the tangent space with translation vector 12 at θ∗.

B. A Simple yet Illustrative Example

The following example illustrates the different notionsof synchronization introduced above and points out variousimportant geometric subtleties occurring on the compact statespace T2. Consider n = 2 oscillators with ω2 ≥ 0 ≥ ω1 =−ω2. We restrict our attention to angles contained in an openhalf-circle: for angles θ1, θ2 with |θ2− θ1| < π, the angular

1For example, in power network applications the coupling termsaij sin(θi − θj) are power flows along transmission lines i, j ∈ E , andthe phase distances |θi − θj | are bounded well below π/2 due to thermalconstraints. In Subsection III-E, we present a converse synchronizationnotion, where the goal is to maximize phase distances.

difference θ2 − θ1 is the number in ]−π, π[ with magnitudeequal to the geodesic distance |θ2 − θ1| and with positivesign if and only if the counter-clockwise path length fromθ1 to θ2 on T1 is smaller than the clockwise path length.With this definition the two-dimensional oscillator dynamics(θ1, θ2) can be reduced to the scalar difference dynamicsθ2−θ1. After scaling time as t 7→ t(ω2−ω1) and introducingκ = 2a12/(ω2 − ω1) the difference dynamics are

d

d t(θ2 − θ1) = fκ(θ2 − θ1) := 1− κ sin(θ2 − θ1) . (11)

The scalar dynamics (11) can be analyzed graphically byplotting the vector field fκ(θ2 − θ1) over the differencevariable θ2 − θ1, as in Figure 5(a). Figure 5(a) displays asaddle-node bifurcation at κ = 1. For κ < 1 no equilibriumof (11) exists, and for κ > 1 an asymptotically stableequilibrium θstable = arcsin(κ−1) ∈ ]0, π/2[ together witha saddle point θsaddle = arcsin(κ−1) ∈ ]π/2, π[ exists.

For θ(0) ∈ Arcn(|θsaddle|) all trajectories converge ex-ponentially to θstable, that is, the oscillators synchronizeexponentially. Additionally, the oscillators are phase cohesiveif an only if θ(0) ∈ Arcn(|θsaddle|), where all trajectoriesremain bounded. For θ(0) 6∈ Arcn(|θsaddle|) the differenceθ2(t) − θ1(t) will increase beyond π, and by definitionwill change its sign since the oscillators change orientation.Ultimately, θ2(t)− θ1(t) converges to the equilibrium θstablein the branch where θ2 − θ1 < 0. In the configuration spaceT2 this implies that the distance |θ2(t)− θ1(t)| increases toits maximum value π and shrinks again, that is, the oscil-lators are not phase cohesive and revolve once around thecircle before converging to the equilibrium manifold. Sincesin(θstable) = sin(θsaddle) = κ−1, strongly coupled oscillatorswith κ 1 practically achieve phase synchronization fromevery initial condition in an open semi-circle. In the criticalcase, κ = 1, the saddle equilibrium manifold at π/2 isglobally attractive but not stable. An representative trajectoryis illustrated in Figure 5(b).

! !"# $ $"# % %"# &!$

!!"#

!

!"#

$

$"#

κ = 1κ > 1

κ < 1

θ2 − θ1

f κ(θ

2−θ 1

)

(a) Vector field (11) for θ2−θ1 > 0 (b) Trajectory θ(t) for κ = 1

Fig. 5. Plot of the vector field (11) for various values of κ and a trajectoryθ(t) ∈ T2 for the critical case κ = 1, where the dashed line is the saddleequilibrium manifold and and • depict θ(0) and limt→∞ θ(t). The non-smoothness of the vector field f(θ2 − θ1) at the boundaries 0, π is anartifact of the non-smoothness of the geodesic distance on T2

In conclusion, the simple but already rich 2-dimensionalcase shows that two oscillators are phase cohesive andsynchronize if and only if κ > 1, that is, if and only if thecoupling dominates the non-uniformity as 2a12 > ω2 − ω1.The ratio 1/κ determines the ultimate phase cohesiveness as

Page 7: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

well as the set of admissible initial conditions. For κ 1,practical phase synchronization is achieved for all anglesin an open semi-circle. More general coupled oscillatornetworks display the same phenomenology, but the thresholdfrom incoherence to synchrony is generally unknown.

C. Synchronization Metrics

The notion of phase cohesiveness can be understood as aperformance measure for synchronization and phase synchro-nization is simply the extreme case of phase cohesivenesswith limt→∞ θ(t) ∈ ∆G(0) = Arcn(0). An alternativeperformance measure is the magnitude of the so-called orderparameter introduced by Kuramoto [5], [6]:

reiψ =1

n

∑n

j=1eiθj .

The order parameter is the centroid of all oscillators repre-sented as points on the unit circle in C1. The magnitude rof the order parameter is a synchronization measure: if alloscillators are phase-synchronized, then r = 1, and if all os-cillators are spaced equally on the unit circle, then r = 0. Thelatter case is characterized in Subsection III-E. For a com-plete graph, the magnitude r of the order parameter serves asan average performance index for synchronization, and phasecohesiveness can be understood as a worst-case performanceindex. Extensions of the order parameter tailored to non-complete graphs have been proposed in [19], [52], [56].

For a complete graph and for γ sufficiently small, the set∆G(γ) reduces to Arcn(γ), the arc of length γ containingall oscillators. The order parameter is contained withinthe convex hull of this arc since it is the centroid of alloscillators, see Figure 6. In this case, the magnitude r of theorder parameter can be related to the arc length γ.

rmin rmax

Fig. 6. Schematic illustration of an arc of length γ ∈ [0, π], its convex hull(shaded), and the value • of the corresponding order parameter reiψ withminimum magnitude rmin = cos(γ/2) and maximum magnitude rmax = 1.

Lemma 3.1: (Shortest arc length and order parameter,[74, Lemma 2.1]) Given an angle array θ = (θ1, . . . , θn) ∈Tn with n ≥ 2, let r(θ) = 1

n |∑nj=1 e

iθj | be the magnitudeof the order parameter, and let γ(θ) be the length of theshortest arc containing all angles, that is, θ ∈ Arcn(γ(θ)).The following statements hold:

1) if γ(θ) ∈ [0, π], then r(θ) ∈ [cos(γ(θ)/2), 1]; and2) if θ ∈ Arcn(π), then γ(θ) ∈ [2 arccos(r(θ)), π].

D. Synchronization Conditions

The coupled oscillator dynamics (1) feature (i) the syn-chronizing coupling described by the graph G(V, E , A) and(ii) the de-synchronizing effect of the non-uniform natural

frequencies ω. Loosely speaking, synchronization occurswhen the coupling dominates the non-uniformity. Variousconditions have been proposed in the synchronization andpower systems literature to quantify this trade-off.

The coupling is typically quantified by the algebraicconnectivity λ2(L) [44], [45], [52], [64], [127], [128] or theweighted nodal degree degi ,

∑nj=1 aij [64], [97], [117],

[129], [130], and the non-uniformity is quantified by eitherabsolute norms ‖ω‖p or incremental norms ‖BTω‖p, wheretypically p ∈ 2,∞. Sometimes, these conditions can beevaluated only numerically since they are state-dependent[127], [129] or arise from a non-trivial linearization process,such as the Master stability function formalism [44], [45],[131]. In general, concise and accurate results are known onlyfor specific topologies such as complete graphs [74], linearchains [108], and bipartite graphs [82] with uniform weights.

For arbitrary coupling topologies only sufficient conditionsare known [52], [64], [127], [129] as well as numericalinvestigations for random networks [89], [98], [99], [128],[132]. Simulation studies indicate that these conditions areconservative estimates on the threshold from incoherence tosynchrony. Literally, every review article on synchronizationdraws attention to the problem of finding sharp synchroniza-tion conditions [7], [8], [44]–[46], [74], [114].

E. Phase Balancing and Splay State

In certain applications in neuroscience [11]–[13], deep-brain stimulation [26], [27], and vehicle coordination [19],[28]–[31], one is not interested in the coherent behaviorwith synchronized (or nearly synchronized) phases, but ratherin the phenomenon of synchronized frequencies and de-sychronized phases.

Whereas the phase-synchronized state is characterized bythe order parameter r achieving its maximal (unit) magni-tude, we say that a solution θ : R≥0 → Tn to the coupledoscillator model (1) achieves phase balancing if all phasesθi(t) converge to Baln = θ ∈ Tn : r(θ) = | 1n

∑nj=1 e

iθj | =0 as t→∞, that is, the oscillators are distributed over theunit circle S1, such that their centroid reiψ vanishes. We referto [28] for a geometric characterization of the balanced state.

One balanced state of particular interest in neuroscienceapplications [11]–[13], [26], [27] is the so-called splay stateθ ∈ Tn : θi = i · 2π/n+ ϕ (mod 2π) , ϕ ∈ S1 , i ∈1, . . . , n ⊆ Baln corresponding to phases uniformlydistributed around the unit circle S1 with distances 2π/n.Other highly symmetric balanced states consist of multipleclusters of collocated phases, where the clusters themselvesare arranged in splay state, see [28], [29].

IV. ANALYSIS OF SYNCHRONIZATION

In this section we present several analysis approaches tosynchronization in the coupled oscillator model (1). We beginwith a few basic ideas to provide important intuition as wellas the analytic basis for further analysis.

Page 8: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

A. Some Simple Yet Important Insights

The potential energy U : Tn → R of the elastic springnetwork in Figure 1 is, up to an additive constant, given by

U(θ) =∑i,j∈E

aij(1− cos(θi − θj)

). (12)

By means of the potential energy, the coupled oscillatormodel (1) can reformulated as the forced gradient system

θi = ωi −∇iU(θ) , i ∈ 1, . . . , n , (13)

where ∇iU(θ) = ∂∂θiU(θ) denotes the partial derivative.

It can be easily verified that the phase-synchronized stateθi = θj for all i, j ∈ E is a local minimum of the potentialenergy (12). The gradient formulation (13) clearly empha-sizes the competition between the synchronization-enforcingcoupling through the potential U(θ) and the synchronization-inhibiting heterogeneous natural frequencies ωi.

We next note that ω has to satisfy certain bounds, relativeto the weighted nodal degree, in order for a synchronizedsolution to exist.

Lemma 4.1: (Necessary sync conditions) Consider thecoupled oscillator model (1) with graph G(V, E , A), fre-quencies ω ∈ 1⊥n , and nodal degree degi =

∑nj=1 aij for

each oscillator i ∈ 1, . . . , n. If there exists a synchronizedsolution θ ∈ ∆G(γ) for some γ ∈ [0, π/2], then thefollowing conditions hold:

1) Absolute bound: For each node i ∈ 1, . . . , n,

degi sin(γ) ≥ |ωi| ; (14)

2) Incremental bound: For all distinct i, j ∈ 1, . . . , n,

(degi + degj) sin(γ) ≥ |ωi − ωj | . (15)

Proof: Since ω ∈ 1⊥n , the synchronization frequencyωsync is zero, and phase and frequency synchronized solutionsare equilibrium solutions determined by the equations

ωi =∑n

j=1aij sin(θi − θj) , i ∈ 1, . . . , n . (16)

Since sin(θi−θj) ∈ [− sin(γ),+ sin(γ)] for θ ∈ ∆G(γ), theequilibrium equations (16) have no solution if condition (14)is not satisfied. Since ω ∈ 1⊥n , an incremental bound on ωseems to be more appropriate than an absolute bound. Thesubtraction of the ith and jth equation (16) yields

ωi − ωj =∑n

k=1(aik sin(θi − θk)− ajk sin(θj − θk)) .

Again, since the coupling is bounded, the above equation hasno solution in ∆G(γ) if condition (15) is not satisfied.

The following result is fundamental for various approachesto phase and frequency synchronization. To the best of theauthors’ knowledge this result has been first established in[133], and it has been reproved numerous times.

Lemma 4.2: (Stable synchronization in ∆G(π/2)) Con-sider the coupled oscillator model (1) with a connectedgraph G(V, E , A) and frequencies ω ∈ 1⊥n . The followingstatements hold:

1) Jacobian: The Jacobian J(θ) of the coupled oscillatormodel (1) evaluated at θ ∈ Tn is given by

J(θ) = −B diag(aij cos(θi − θj)i,j∈E)BT ;

2) Local stability and uniqueness: If there exists anequilibrium θ∗ ∈ ∆G(π/2), then

(i) −J(θ∗) is a Laplacian matrix;(ii) the equilibrium manifold [θ∗] ∈ ∆G(π/2) is

locally exponentially stable; and(iii) this equilibrium manifold is unique in ∆G(π/2).

Proof: Since ∂∂θi

(ωi −

∑nj=1 aij sin(θi − θj)

)=

−∑nj=1 aij cos(θi − θj) and ∂

∂θj

(ωi −

∑nj=1 aij sin(θi −

θj))

= aij cos(θi − θj), we obtain that the Jacobian isequal to minus the Laplacian matrix of the connected graphG(V, E , A) with the (possibly negative) weights aij =aij cos(θi − θj). Equivalently, in compact notation J(θ) =−B diag(aij cos(θi − θj)i,j∈E)BT . This completes theproof of statement 1).

The Jacobian J(θ) evaluated for an equilibrium θ∗ ∈∆G(π/2) is minus the Laplacian matrix of the graphG(V, E , A) with strictly positive weights aij = aij cos(θ∗i −θ∗j ) > 0 for every i, j ∈ E . Hence, J(θ∗) is negativesemidefinite with the nullspace 1n arising from the rotationalsymmetry, see Figure 4. Consequently, the equilibrium pointθ∗ ∈ ∆G(π/2) is locally (transversally) exponentially sta-ble, or equivalently, the corresponding equilibrium manifold[θ∗] ∈ ∆G(π/2) is locally exponentially stable.

The uniqueness statement follows since the right-hand sideof the coupled oscillator model (1) is a one-to-one function(modulo rotational symmetry) for θ ∈ ∆G(π/2), see [134,Corollary 1]. This completes the proof of statement 2).

By Lemma 4.2, any equilibrium in ∆G(π/2) is stablewhich supports the notion of phase cohesiveness as a per-formance metric. Since the Jacobian J(θ) is the negativeHessian of the potential U(θ) defined in (12), Lemma 4.2also implies that any equilibrium in ∆G(π/2) is a localminimizer of U(θ). Of particular interest are so-called S1-synchronizing graphs for which all critical points of (12)are hyperbolic, the phase-synchronized state is the globalminimum of U(θ), and all other critical points are localmaxima or saddle points. The class of S1-synchronizinggraphs includes, among others, complete graphs and acyclicgraphs [100]–[103].

These basic insights motivated various characterizationsand explorations of the critical points and the curvature ofthe potential U(θ) in the literature on synchronization [52],[64], [74], [89], [93], [100], [100]–[103], [103] as well ason power systems [61], [116], [127], [129], [133]–[137].

B. Phase Synchronization

If all natural frequencies are identical, ωi ≡ ω for all i ∈1, . . . , n, then a transformation of the coupled oscillatormodel (1) to a rotating frame with frequency ω leads to

θi = −∑n

j=1aij sin(θi − θj) , i ∈ 1, . . . , n . (17)

Page 9: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

The analysis of the coupled oscillator model (17) is par-ticularly simple and local phase synchronization can beconcluded by various analysis methods. A sample of dif-ferent analysis schemes (by far not complete) includesthe contraction property [54], [64], [92], [100], [138],quadratic Lyapunov functions [52], [64], linearization [81],[103], or order parameter and potential function arguments[28], [56], [80].

The following theorem on phase synchronization summa-rizes a collection of results originally presented in [28], [54],[56], [74], [100], [103], and it can be easily proved given theinsights developed in Subsection IV-A.

Theorem 4.3: (Phase synchronization) Consider the cou-pled oscillator model (1) with a connected graph G(V, E , A)and with frequency ω ∈ Rn (not necessarily zero mean). Thefollowing statements are equivalent:

(i) Stable phase sync: there exists a locally exponentiallystable phase-synchronized solution θ ∈ Arcn(0) (or asynchronization manifold [θ] ∈ ∆G(0)); and

(ii) Uniformity: there exists a constant ω ∈ R such thatωi = ω for all i ∈ 1, . . . , n.

If the two equivalent cases (i) and (ii) are true, the followingstatements hold:

1) Global convergence: For all initial angles θ(0) ∈ Tnall frequencies θi(t) converge to ω and all phasesθi(t) − ωt (mod 2π) converge to the critical pointsθ ∈ Tn : ∇U(θ) = 0n;

2) Semi-global stability: The region of attraction of thephase-synchronized solution θ ∈ Arcn(0) contains theopen semi-circle Arcn(π), and each arc Arcn(γ) ispositively invariant for every arc length γ < π;

3) Explicit phase: For initial angles in an open semi-circle θ(0) ∈ Arcn(π), the asymptotic synchronizationphase is given by2 θ(t)=

∑ni=1θi(0)/n+ωt (mod 2π);

4) Convergence rate: For every initial angle θ(0) ∈Arcn(γ) with γ < π, the exponential convergencerate to phase synchronization is no worse than λps =−λ2(L) sinc(γ); and

5) Almost global stability: If the graph G(V, E , A) isS1-synchronizing, the region of attraction of the phase-synchronized solution θ ∈ Arcn(0) is almost all of Tn.

Proof: Implication (i) =⇒ (ii): By assumption, thereexist constants θsync ∈ S1 and ωsync ∈ R such that θi(t) =θsync +ωsynct (mod 2π). In the phase-synchronized case, thedynamics (1) then read as ωsync = ωi for all i ∈ 1, . . . , n.Hence, a necessary condition for the existence of phasesynchronization is that all ωi are identical.

Implication (ii) =⇒ (i): Consider the model (1) writtenin a rotating frame with frequency ω as in (17). Note that theset of phase-synchronized solutions ∆G(0) is an equilibriummanifold. By Lemma 4.2, we conclude that ∆G(0) is locallyexponentially stable. This concludes the proof of (i) ⇔ (ii).

2This “average” of angles (points on S1) is well-defined in an opensemi-circle. If the parametrization of θ has no discontinuity inside the arccontaining all angles, then the average can be obtained by the usual formula.

Statement 1): Note that (17) can be written as the gradientflow θ = −∇U(θ), and the corresponding potential functionU(θ) is non-increasing along trajectories. Since the sublevelsets of U(θ) are compact and the vector field ∇U(θ) issmooth, the invariance principle [139, Theorem 4.4] assertsthat every solution converges to set of equilibria of (17).

Statements 2): The coupled oscillator model (17) can bere-written as the consensus-type system

θi = −∑n

j=1bij(θ) · (θi − θj) , i ∈ 1, . . . , n , (18)

where the weights bij(θ) = aij sinc(θi − θj) depend ex-plicitly on the system state. Notice that for θ ∈ Arcn(γ)and γ < π the weights bij(θ) are upper and lower boundedas bij(θ) ∈ [aij sinc(γ), aij ] Assume that the initial anglesθi(0) belong to the set Arcn(γ), that is, they are all containedin an arc of length γ ∈ [0, π[. In this case, a natural Lyapunovfunction to establish phase synchronization can be obtainedfrom the contraction property, which aims at showing thatthe convex hull containing all oscillators is decreasing, see[54], [64], [92], [100], [140] and the review [138, Section 2].

Recall the geodesic distance between two angles on S1and define the continuous function V : Tn → [0, π] by

V (ψ) = max|ψi − ψj | | i, j ∈ 1, . . . , n. (19)

Notice that, if all angles are contained in an arc at time t,then the arc length V (θ(t)) = maxi,j∈1,...,n |θi(t)− θj(t)|is a Lyapunov function candidate for phase synchronization.Indeed, it can be shown that V (θ(t)) decreases along trajec-tories of (18) for θ(0) ∈ Arcn(γ) and for all γ < π. Theanalysis is complicated by the following fact: the functionV (θ(t)) is continuous but not necessarily differentiable whenthe maximum geodesic distance (that is, the right-hand sideof (19)), is attained by more than one pair of oscillators. Weomit the explicit calculations here and refer to [54], [64],[74], [83], [92] for a detailed analysis.

Statement 3): By statement 2), the set Arcn(π) is pos-itively invariant, and for θ(0) ∈ Arcn(π) the average∑ni=1 θi(t)/n is well defined for t ≥ 0. A summation

over all equations of the model (17) yields∑ni=1 θi(t) =

0, or equivalently,∑ni=1 θi(t) is constant for all t ≥ 0.

In particular, for t = 0 we have that∑ni=1 θi(t) =∑n

i=1 θi(0) and for a phase-synchronized solution we havethat

∑ni=1 θsync =

∑ni=1 θi(0). Hence, the explicit synchro-

nization phase is given by∑ni=1 θi(0)/n. In the original

coordinates (non-rotating frame) the synchronization phaseis given by

∑ni=1 θi(0)/n+ ωt (mod 2π).

Statement 4): Given the invariance of the set Arcn(γ) forany γ < π, the system (18) can be analyzed as a lineartime-varying consensus system with initial condition θ(0) ∈Arcn(γ), and bounded time-varying weights bij(θ(t)) ∈[aij sinc(γ), aij ] for all t ≥ 0. The worst-case convergencerate λps can then be obtained by a standard symmetricconsensus analysis, see [52], [53], [64], [74]. For instance,it can be shown that the deviation of the angles θ(t) fromtheir average, ‖θ(t)− (

∑ni=1 θ(t)/n)1n‖22 (the disagreement

function) decays exponentially with rate λps.

Page 10: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Statement 5): By statement 1), all solutions of system (17)converge to the set of equilibria, which equals the set ofcritical points of the potential U(θ). By the definition of S1-synchronizing graphs, the phase-synchronized equilibriummanifold Arcn(0) is the only stable equilibrium set, and allothers are unstable. Hence, for all initial condition θ(0) ∈Tn, which are not on the stable manifolds of unstableequilibria, the corresponding solution θ(t) will reach thephase-synchronized equilibrium manifold Arcn(0).

Remark 2: (Control-theoretic perspective on synchro-nization) As established in Theorem 4.3, the set of phase-synchronized solutions Arcn(0) of the coupled oscillatormodel (1) is locally stable provided that all natural frequen-cies are identical. For non-uniform (but sufficiently identical)natural frequencies, phase synchronization is not possible buta certain degree of phase cohesiveness can still be achieved.Hence, the coupled oscillator model (1) can be regarded asan exponentially stable system subject to the disturbanceω ∈ 1⊥n , and classic control-theoretic concepts such as input-to-state stability, practical stability, and ultimate boundedness[139] or their incremental versions [141] can be used to studysynchronization. In control-theoretic terminology, synchro-nization and phase cohesiveness can then also be described as“practical phase synchronization”. Compared to prototypicalnonlinear control examples, various additional challengesarise in the analysis of the coupled oscillator model (1) due tothe bounded and non-monotone sinusoidal coupling and thecompact state space Tn containing numerous equilibria; seethe analysis approaches in Section IV and [64], [74], [95].

C. Phase Balancing

In general, only few results are known about the phasebalancing problem. This asymmetry is partially caused bythe fact that phase synchrony is required in more applicationsthan phase balancing. Moreover, the phase-synchronized setArcn(0) admits a very simple geometric characterization,whereas the phase-balanced set Baln has a complicated struc-ture consisting of numerous disjoint subsets. The numberof these subsets grows with the number of nodes n in acombinatorial fashion.

Consider the coupled oscillator model (17) with identicalnatural frequencies. By inverting the direction of time, we get

θi =∑n

j=1aij sin(θi − θj) , i ∈ 1, . . . , n . (20)

In the following, we say that the interaction graph G(V, E , A)is circulant if the adjacency matrix A = AT is a circulantmatrix. Circulant graphs are highly symmetric graphs includ-ing complete graphs, bipartite graphs, and ring graphs.3 Forcirculant and uniformly weighted graphs, the coupled oscil-lator model (20) achieves phase balancing. The followingtheorem summarizes different results, which were originallypresented in [28], [29].

Theorem 4.4: (Phase balancing) Consider the coupledoscillator model (20) with a connected, uniformly weighted,

3Further info on circulant graphs and a gallery can be found athttp://mathworld.wolfram.com/CirculantGraph.html.

and circulant graph G(V, E , A). The following statementshold:

1) Local phase balancing: The phase-balanced set Balnis locally asymptotically stable; and

2) Almost global stability: If the graph G(V, E , A) iscomplete, then the region of attraction of the stablephase-balanced set Baln is almost all of Tn.

The proof of Theorem 4.4 follows a similar reasoningas the proof of Theorem 4.3: convergence is establishedby potential function arguments and local (in)stability ofequilibria by Jacobian arguments. We omit the proof hereand refer to [28, Theorem 1] and [29, Theorem 2] for details.

For general connected graphs, the conclusions of Theorem4.4 are not true. As a remedy to achieve locally stable andglobally attractive phase balancing, higher order models needto be considered, see the models proposed in [29], [56].

D. Synchronization in Complete Networks

For a complete coupling graph with uniform weights aij =K/n, where K > 0 is the coupling gain, the coupled oscil-lator model (1) reduces to the celebrated Kuramoto model

θi = ωi−K

n

∑n

j=1sin(θi−θj) , i ∈ 1, . . . , n . (21)

By means of the order parameter reiψ = 1n

∑nj=1 e

iθj , theKuramoto model (21) can be rewritten in the insightful form

θi = ωi −Kr sin(θi − ψ) , i ∈ 1, . . . , n . (22)

Equation (22) gives the intuition that the oscillators syn-chronize by coupling to a mean field represented by theorder parameter reiψ . Intuitively, for small coupling strengthK each oscillator rotates with its natural frequency ωi,whereas for large coupling strength K all angles θi(t) willbe entrained by the mean field reiψ and the oscillatorssynchronize. The threshold from incoherence to synchroniza-tion occurs for some critical coupling Kcritical. This phasetransition has been the source of numerous investigationsstarting with Kuramoto’s analysis [5], [6]. Various necessary,sufficient, implicit, and explicit estimates of the criticalcoupling strength Kcritical for both the on-set as well asthe ultimate stage of synchronization have been proposed[5]–[8], [28], [52], [53], [64], [74], [75], [82]–[87], [95]–[97], [100], [103]–[107], [110], and we refer to [74] for acomprehensive overview.

The mean field approach to the equations (22) can be mademathematically rigorous by a time-scale separation [96] orin the continuum limit as the number of oscillators tendsto infinity and the natural frequencies ω are sampled froma distribution function g : R → R≥0. In the continuumlimit and for a symmetric, continuous, and unimodal dis-tribution g(ω), Kuramoto himself showed in an insightfuland ingenuous analysis [5], [6] that the incoherent state (auniform distribution of the oscillators on the unit circle S1)supercritically bifurcates for the critical coupling strength

Kcritical =2

πg(0). (23)

Page 11: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

In [8], [87], [104], it was found that the bipolar (bimodaldouble-delta) distribution (respectively the uniform distri-bution) yield the largest (respectively smallest) thresholdKcritical over all distributions g(ω) with bounded support. Werefer [7], [8] for further references and to [88], [109]–[111]for recent contributions on the continuum limit model.

In the finite-dimensional case, the necessary synchroniza-tion condition (15) gives a lower bound for Kcritical as

K ≥ n

2(n− 1)· (ωmax − ωmin) . (24)

Three recent articles [84]–[86] independently derived aset of implicit consistency equations for the exact criticalcoupling strength Kcritical for which synchronized solutionsexist. Verwoerd and Mason provided the following implicitformulae to compute Kcritical [86, Theorem 3]:

Kcritical = nu∗/∑n

i=1

√1− (Ωi/u∗)2 ,

2∑n

i=1

√1− (Ωi/u∗)2 =

∑n

i=11/√

1− (Ωi/u∗)2,(25)

where Ωi = ωi − ωsync and u∗ ∈ [‖Ω‖∞ , 2 ‖Ω‖∞]. The im-plicit formulae (25) can also be extended to bipartite graphs[82]. A local stability analysis is carried out in [84], [85].

From the point of analyzing or designing a sufficientlystrong coupling, the exact formulae (25) have three draw-backs. First, they are implicit and thus not suited for perfor-mance or robustness estimates in case of additional couplingstrength for a given K > Kcritical. Second, the correspondingregion of attraction of a synchronized solution is unknown.Third and finally, the particular natural frequencies ωi aretypically time-varying, uncertain, or even unknown in theapplications listed in Section I. In this case, the exact valueof Kcritical needs to be estimated in continuous time, or aconservatively strong coupling KKcritical has to be chosen.

The following theorem states an explicit bound on the crit-ical coupling strength together with performance estimates,convergence rates, and a guaranteed semi-global region ofattraction for synchronization. This bound is tight and thusnecessary and sufficient when considering arbitrary distribu-tions of the natural frequencies with compact support. Theresult has been originally presented in [74, Theorem 4.1].

Theorem 4.5: (Synchronization in the Kuramotomodel) Consider the Kuramoto model (21) with naturalfrequencies ω = (ω1, . . . , ωn) and coupling strength K. Thefollowing three statements are equivalent:

(i) the coupling strength K is larger than the maximumnon-uniformity among the natural frequencies, that is,

K > Kcritical , ωmax − ωmin ; (26)

(ii) there exists an arc length γmax ∈ ]π/2, π] such that theKuramoto model (21) synchronizes exponentially forall possible distributions of the natural frequencies ωisupported on the compact interval [ωmin, ωmax] and forall initial phases θ(0) ∈ Arcn(γmax); and

(iii) there exists an arc length γmin ∈ [0, π/2[ such that theKuramoto model (21) has a locally exponentially stablesynchronization manifold in Arcn(γmin) for all possible

distributions of the natural frequencies ωi supported onthe compact interval [ωmin, ωmax].

If the three equivalent conditions (i), (ii), and (iii) hold,then the ratio Kcritical/K and the arc lengths γmin ∈ [0, π/2[and γmax ∈ ]π/2, π] are related uniquely via sin(γmin) =sin(γmax) = Kcritical/K, and the following statements hold:

1) phase cohesiveness: the set Arcn(γ) ⊆ ∆G(γ) ispositively invariant for every γ ∈ [γmin, γmax], and eachtrajectory starting in Arcn(γmax) approaches asymptot-ically Arcn(γmin);

2) frequency synchronization: the asymptotic synchro-nization frequency is the average frequency ωsync =1n

∑ni=1 ωi, and, given phase cohesiveness in Arcn(γ)

for some fixed γ < π/2, the exponential synchroniza-tion rate is no worse than λK = −K cos(γ); and

3) order parameter: the asymptotic value of the mag-nitude of the order parameter, denoted by r∞ ,limt→∞ 1

n |∑nj=1 e

iθj(t)|, is bounded as

1 ≥ r∞ ≥ cos(γmin

2

)=

√1 +

√1− (Kcritical/K)2

2.

Proof: In the following, we sketch the proof of Theo-rem 4.5 and refer to [74, Theorem 4.1] for further details.

Implication (i) =⇒ (ii): In a first step, it is shown that thephase cohesive set Arcn(γ) is positively invariant for everyγ ∈ [γmin, γmax]. By assumption, the angles θi(t) belong tothe set Arcn(γ) at time t = 0, that is, they are all containedin an arc of length γ. We aim to show that all angles remainin Arcn(γ) for all subsequent times t > 0 by means of thecontraction Lyapunov function (19). Note that Arcn(γ) ispositively invariant if and only if V (θ(t)) does not increase atany time t such that V (θ(t)) = γ. The upper Dini derivativeof V (θ(t)) along trajectories of (21) is given by

D+V (θ(t)) = limh↓0

supV (θ(t+ h))− V (θ(t))

h.

Written out in components and after trigonometric simplifi-cations [74], we obtain that the derivative is bounded as

D+V (θ(t)) ≤ ωmax − ωmin −K sin(γ) .

It follows that the length of the arc formed by the angles isnon-increasing in Arcn(γ) if and only if

K sin(γ) ≥ Kcritical , (27)

where Kcritical is as stated in equation (26). For γ ∈ [0, π] theleft-hand side of (27) is a concave function of γ that achievesits maximum at γ∗ = π/2. Therefore, there exists an open setof arc lengths γ ∈ [0, π] satisfying equation (27) if and onlyif equation (27) is true with the strict equality sign at γ∗ =π/2, which corresponds to condition (26). Additionally, ifthese two equivalent statements are true, then there existsa unique γmin ∈ [0, π/2[ and a γmax ∈ ]π/2, π] that satisfyequation (27) with the equality sign, namely sin(γmin) =sin(γmax) = Kcritical/K. For every γ ∈ [γmin, γmax] it followsthat the arc-length V (θ(t)) is non-increasing, and it is strictlydecreasing for γ ∈ ]γmin, γmax[. Among other things, this

Page 12: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

shows that statement (i) implies statement 1). By means ofLemma 3.1, statement 3) then follows from statement 1).

The frequency dynamics of the Kuramoto model (21) canbe obtained by differentiating the Kuramoto model (21) as

d

d tθi =

∑n

j=1aij(t) (θj − θi) , (28)

where aij(t) = (K/n) cos(θi(t) − θj(t)). For K > Kcritical,we just proved that for every θ(0) ∈ Arcn(γmax) and for allγ ∈ ]γmin, γmax[ there exists a finite time T ≥ 0 such thatθ(t) ∈ Arcn(γ) for all t ≥ T . Consequently, the terms aij(t)are strictly positive for all t ≥ T . Notice also that system(28) evolves on the tangent space of Tn, that is, the Euclideanspace Rn. Now fix γ ∈ ]γmin, π/2[ and let T ≥ 0 such thataij(t) > 0 for all t ≥ T . In this case, the frequency dynamics(28) can be analyzed as linear time-varying consensus sys-tem. Consider the disagreement vector x = θ − ωsync1n asan error coordinate. By standard consensus arguments [48]–[50], it can be shown that the disagreement vector satisfies‖x(t)‖ ≤ ‖x(0)‖e−λKt for all t ≥ T . This proves statement2) and the implication (i) =⇒ (ii).

Implication (ii) =⇒ (i): To show that condition (26) isalso necessary for synchronization, it suffices to construct acounter example for which K ≤ Kcritical and the oscillatorsdo not achieve exponential synchronization even though allωi ∈ [ωmin, ωmax] and θ(0) ∈ Arcn(γ) for every γ ∈ ]π/2, π].A basic instability mechanism under which synchronizationbreaks down is caused by a bipolar distribution of the naturalfrequencies. Let the index set 1, . . . , n be partitioned bythe two non-empty sets I1 and I2. Let ωi = ωmin for i ∈ I1and ωi = ωmax for i ∈ I2, and assume that at some timet ≥ 0 it holds that θi(t) = −γ/2 for i ∈ I1 and θi(t) =+γ/2 for i ∈ I2 and for some γ ∈ [0, π[. By construction,at time t all oscillators are contained in an arc of lengthγ ∈ [0, π[. Assume now that K<Kcritical and the oscillatorssynchronize. It can be shown [74] that the evolution of thearc length V (θ(t)) satisfies the equality

D+V (θ(t)) = ωmax − ωmin −K sin(γ) . (29)

Clearly, for K < Kcritical the arc length V (θ(t)) = γ isincreasing for any arbitrary γ ∈ [0, π]. Thus, the phases arenot bounded in Arcn(γ). This contradicts the assumptionthat the oscillators synchronize for K < Kcritical from everyinitial condition θ(0) ∈ Arcn(γ). For K = Kcritical, weknow from [84], [85] that phase-locked equilibria have azero eigenvalue with a two-dimensional Jacobian block, andthus synchronization cannot occur. This instability via atwo-dimensional Jordan block is also visible in (29) sinceD+V (θ(t)) is increasing for θ(t) ∈ Arcn(γ), γ ∈ ]π/2, π]until all oscillators change orientation, just as in the examplein Subsection III-B. This proves the implication (ii) =⇒ (i).

Equivalence (i),(ii) ⇔ (iii): The proof relies on Jacobianarguments and will be omitted here, see [74] for details.

Theorem 4.5 places a hard bound on the critical couplingstrength Kcritical for all distributions of ωi supported on thecompact interval [ωmin, ωmax]. For a particular distribution

g(ω) supported on [ωmin, ωmax] the bound (26) is only suffi-cient and possibly a factor 2 larger than the necessary bound(24). The exact critical coupling lies somewhere in betweenand can be obtained from the implicit equations (25).

Since the bound (26) on Kcritical is exact [74] for the worst-case bipolar distribution ωi ∈ ωmin, ωmax, Figure 7 reportsnumerical findings for the other extreme case [87] of a uni-form distribution g(ω) = 1/2 supported for ωi ∈ [−1, 1]. All

n

4/π

Kcr

itic

al

Fig. 7. Statistical analysis of the necessary and explicit bound (24) (♦),the exact and implicit bound (25) (), and the sufficient, tight, and explicitbound (26) () for n ∈ [2, 300] oscillators in a semi-log plot, where thecoupling gains for each n are averaged over 1000 simulations.

three displayed bounds are identical for n = 2 oscillators. Asn increases, the sufficient bound (26) converges to the widthωmax − ωmin = 2 of the support of g(ω), and the necessarybound (24) accordingly to half of the width. The exact bound(25) converges to 4(ωmax − ωmin)/(2π) = 4/π in agreementwith condition (23) predicted for the continuum limit.

Finally, let us mention that Theorem 4.5 can be extendedto discontinuously switching and slowly time-varying naturalfrequencies [74]. For a particular sampling distribution g(ω),the critical quantity in condition (26), the support ωmax−ωmin,can be estimated by extreme value statistics, see [89].

E. Synchronization in Sparse Networks

As summarized in Subsection III-D, the quest for sharpand concise synchronization for non-complete couplinggraph G(V, E , A) is an important and outstanding problememphasized in every review article on coupled oscillatornetworks [7], [8], [44]–[46], [74]. The approaches known forphase synchronization in arbitrary graphs or the contractionapproach to frequency synchronization (used in the proof ofTheorem 4.5) do not generally extend to arbitrary natural fre-quencies ω ∈ 1⊥n and connected coupling graphs G(V, E , A),or do so only under extremely conservative conditions.

One Lyapunov function advocated for classic Kuramotooscillators (21) is the function W : Arcn(π) → R definedfor angles θ in an open semi-circle and given by [52], [53]

W (θ) =1

4

∑n

i,j=1|θi − θj |2 =

1

2

∥∥BTc θ∥∥22 , (30)

where Bc ∈ Rn×(n(n−1)/2) is an incidence matrix of thecomplete graph. As shown in [64, Theorem 4.4], the Lya-punov function (30) generalizes also to the coupled oscillatormodel (1). Indeed, an even more general model is consideredin [64], and a Lyapunov analysis yields the following result.

Page 13: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Theorem 4.6: (Frequency synchronization I) Considerthe coupled oscillator model (1) with a connected graphG(V, E , A) and ω ∈ 1⊥n . Assume that the algebraic con-nectivity is larger than a critical value, that is,

λ2(L) > λcritical ,∥∥BTc ω∥∥2 , (31)

where Bc ∈ Rn×n(n−1)/2 is the incidence matrix of thecomplete graph. Accordingly, define γmax ∈ ]π/2, π] andγmin ∈ [0, π/2[ as unique solutions to (π/2) · sinc(γmax) =sin(γmin)=λcritical/λ2(L). The following statements hold:

1) phase cohesiveness: the setθ ∈ Arcn(π) :

‖BTc θ‖2 ≤ γ⊆ ∆G(γ) is positively invariant for ev-

ery γ ∈ [γmin, γmax], and each trajectory starting in thesetθ ∈ Arcn(π) :

∥∥BTc θ∥∥2 < γmax

asymptoticallyreaches the set

θ ∈ Arcn(π) : ‖BTc θ‖2 ≤ γmin

; and

2) frequency synchronization: for every θ(0) ∈ Arcn(π)with

∥∥BTc θ(0)∥∥2< γmax the frequencies θi(t) synchro-

nize exponentially to the average frequency ωsync =1n

∑ni=1 ωi, and, given phase cohesiveness in ∆G(γ)

for some fixed γ < π/2, the exponential synchroniza-tion rate is no worse than λfe = −λ2(L) cos(γ).

The proof of Theorem 4.6 follows a similar ultimate-boundedness strategy as the proof of Theorem 4.5 by usingthe Lyapunov function (30). It can be found in Appendix B.

For classic Kuramoto oscillators (21), condition (31) re-duces to K >

∥∥BTc ω∥∥2. Clearly, the condition K>∥∥BTc ω∥∥2

is more conservative than the bound (26) which reads as K >‖BTc ω‖∞ = ωmax − ωmin. One reason for this conservatismis that the analysis leading to condition (31) requires allphase distances |θi − θj | to be bounded, whereas accordingto Lemma 4.2 only pairwise phase distances |θi − θj |,i, j ∈ E , need to be bounded for stable synchronization.The following result exploits these weaker assumptions andstates a sharper (but only local) synchronization condition.

Theorem 4.7: (Frequency synchronization II) Considerthe coupled oscillator model (1) with a connected graphG(V, E , A) and ω ∈ 1⊥n . There exists a locally exponentiallystable equilibrium manifold [θ] ∈ ∆G(π/2) if

λ2(L) >∥∥BTω∥∥

2. (32)

Moreover, if condition (32) holds, then [θ] is phase cohesivein θ ∈ Tn : ‖BT θ‖2 ≤ γmin ⊆ ∆G(γmin), where γmin ∈[0, π/2[ satisfies sin(γmin) = ‖BTω‖2/λ2(L).

The strategy to prove Theorem 4.7 is inspired by theingenuous analysis in [52, Section IIV.B]. It relies on theinsight gained from Lemma 4.2 that any synchronizationmanifold [θ] ∈ ∆G(π/2) is locally stable, and it formulatesthe existence of such a synchronization manifold as a fixedpoint problem. Here, we follow the basic proof strategy in[52], but we provide a more accurate result together with aself-contained proof which is reported in Appendix C.

V. CONCLUSIONS AND OPEN RESEARCH DIRECTIONS

In this paper we introduced the reader to the coupled os-cillator model (1), we reviewed several applications, we dis-cussed different synchronization notions, and we presented

different analysis approaches to phase synchronization, phasebalancing, and frequency synchronization.

Despite the vast literature, the countless applications,and the numerous theoretic results on the synchronizationproperties of model (1), many interesting and importantproblems are still open. In the following, we summarizelimitations of the existing analysis approaches and presenta few worthwhile directions for future research.

First, in many applications the coupling between theoscillators is not purely sinusoidal. For instance, phase delaysin neuroscience [13], time delays in sensor networks [37],or transfer conductances in power networks [63] lead to a“shifted coupling” of the form sin(θi − θj − ϕij), whereϕij ∈ [−π/2, π/2]. In this case and also for other “skewed”or “symmetry-breaking” coupling functions, many of thepresented analysis schemes either fail or lead to overlyconservative results. Another interesting class of oscillatornetworks are systems of pulse-coupled oscillators featuringhybrid dynamics: impulsive coupling at discrete time instantsand uncoupled continuous dynamics otherwise. This class ofoscillator networks displays a very interesting phenomenol-ogy. For instance, the behavior of identical oscillators cou-pled in a complete graph strongly depends on the curvatureof the uncoupled dynamics [142]. Most of the results knownfor continuously-coupled oscillators still need to be extendedto pulse-coupled oscillators with hybrid dynamics.

Second, in many applications [12], [24], [34], [63], [67]the coupled oscillator dynamics are not given by a simplefirst-order phase model of the form (1). Rather, the dynamicsare of higher order, or sometimes there is no readily availablephase variable to describe the limit cycle attracting the cou-pled dynamics. The analysis of oscillator networks with moregeneral oscillator dynamics is largely unexplored. Whereasadvances have been made for the simple case of phasesynchronization of linear or passive oscillator networks, thecase of frequency synchronization of non-identical oscillatorswith higher-order dynamics is not well-studied.

Third, despite the vast scientific interest the quest forsharp, concise, and closed-form synchronization conditionsfor arbitrary complex graphs has been so far in vain [7],[8], [44]–[46], [74]. As suggested by Lemma 4.1, Lemma4.2, Theorem 4.5, and the proof of Theorem 4.7, the propermetric for the synchronization problem is the incremental∞-norm ‖BT θ‖∞ = maxi,j∈E |θi − θj |. In the authors’opinion, a Banach space analysis of the coupled oscil-lator model (1) with the incremental ∞-norm will mostlikely deliver the sharpest possible conditions. However,such an analysis is very challenging for arbitrary naturalfrequencies ω ∈ 1⊥n and connected and weighted couplinggraphsG(V, E , A). Recent work [114] by the authors putsforth a novel algebraic condition for synchronization witha rigorous analysis for specific classes of graphs and with(only) a statistical validation for generic weighted graphs.

Fourth and finally, a few interesting and open theoreticalchallenges include the following. First, most of the presentedanalysis approaches and conditions do not extend to time-varying or directed coupling graphs G(V, E , A), and alterna-

Page 14: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

tive methods need to be developed. Second, most known es-timates on the region of attraction of a synchronized solutionare conservative. The semi-circle estimates given in Theorem4.3 and Theorem 4.5 rely on convexity of Arcn(π) and areoverly conservative. We refer to [63], [112] for a set ofinteresting results and conjectures on the region of attraction.Third, the presented analysis approaches are restricted tosynchronized equilibria inside the set ∆G(π/2). Other inter-esting equilibrium configurations outside ∆G(π/2) includesplay state equilibria or frequency-synchronized equilibriawith phases spread over an entire semi-circle.

We sincerely hope that this tutorial article stimulatesfurther exciting research on synchronization in coupled oscil-lators, both on the theoretical side as well as in the countlessapplications.

APPENDIX

A. Modeling of the spring-interconnected particles

Consider the spring network in Figure 1 consisting of agroup of particles constrained to rotate around a circle ofunit radius. For simplicity, we assume that the particles areallowed to move freely on the circle and exchange their orderwithout collisions. Each particle is characterized by its phaseangle θi ∈ S1 and frequency θi ∈ R, and its inertial anddamping coefficients are Mi > 0 and Di > 0.

The external forces and torques acting on each particleare (i) a viscous damping force Diθi opposing the directionof motion, (ii) a non-conservative force ωi ∈ R along thedirection of motion depicting a preferred natural rotationfrequency, and (iii) an elastic restoring torque between in-teracting particles i and j coupled by an ideal elastic springwith stiffness aij > 0 and zero rest length.

To compute the elastic torque between the particles, weparametrize the position of each particle i by the unit vectorpi = [cos(θi) , sin(θi)]

T ∈ S1 ⊂ R2. The elastic Hookeanenergy stored in the springs is the function E : Tn → Rgiven up to an additive constant by

E(θ) =∑i,j∈E

aij2‖pi − pj‖22

=∑i,j∈E

aij(1− cos(θi) cos(θj)− sin(θi) sin(θj)

)=∑i,j∈E

aij(1− cos(θi − θj)

),

where we employed the trigonometric identity cos(α−β) =cosα cosβ+sinα sinβ in the last equality. Hence, we obtainthe restoring torque acting on particle i as

Ti(θ) = − ∂

∂θiE(θ) = −

∑i,j∈E

aij sin(θi − θj) .

Therefore, the network of spring-interconnected particlesdepicted in Figure 1 obeys the dynamics

Miθi +Diθi = ωi −∑i,j∈E

aij sin(θi − θj) ,i ∈ 1, . . . , n . (33)

The coupled oscillator model (1) is then obtained as thekinematic variant or the overdamped limit of the spring

network (33) with zero inertia Mi = 0 and unit dampingDi = 1 for all oscillators i ∈ 1, . . . , n.B. Proof of Theorem 4.6

Assume that θ(0) ∈ Arcn(ρ) for ρ ∈ [0, π[. Recall that theangular differences are well defined for θ in the open semi-circle Arcn(π), and define the vector of phase differencesδ , BTc θ = (θ2 − θ1, . . . ) ∈ [−π,+π]

n(n−1)/2. By takingthe derivative d/dt δ(t) the phase differences satisfy

δ = BTc ω −BTc B diag(aiji,j∈E) sin(BT θ)

= BTc ω −BTc Bc diag(aiji,j∈1,...,n,i<j) sin(δ), (34)

where sin(x) = (sin(x1), . . . , sin(xn)) for a vector x ∈ Rn.Notice that for θ(0) ∈ Arcn(π) the δ-dynamics (34) are well-defined for an open interval of time. In the following, wewill show that the set δ ∈ Rn : ‖δ‖2 < γmax is positivelyinvariant under condition (31). As a consequence, the setδ ∈ Rn : ‖δ‖∞ < γmax ≤ π is positively invariant aswell, and the δ-coordinates are well defined for all t ≥ 0.

The Lyapunov function (30) reads in δ-coordinates asW (δ) = 1

2‖δ‖2, and its derivative along trajectories of (34) is

W (δ) = δTBTc ω − δTBTc Bc diag(aiji<j) sin(δ)

= δTBTc ω − n δT diag(aiji<j) sin(δ) , (35)

where the second equality follows from the identity

δTBTc Bc=θTBcBTc Bc=θT (nIn−1n×n)Bc=nθTBc=nδ.

For ‖δ2‖ ≤ ρ, ρ ∈ [0, π[, consider the following inequalities

n δT diag(aiji<j) sin(δ)

= n (BTc θ)T diag(aij sinc(θi − θj)i<j)(BTc θ)

≥ n sinc(ρ) (BTc θ)T diag(aiji<j)(BTc θ)

≥ λ2(L) sinc(ρ)‖BTc θ‖22 = λ2(L) sinc(ρ)‖δ‖22 ,where the last inequality follows from [64, Lemma 4.7].Hence, the derivative (35) simplifies further to

W (δ) ≤ δTBTc ω − λ2(L) sinc(ρ)‖δ‖22. (36)

In the following we regard BTc ω as external disturbanceaffecting the otherwise stable δ-dynamics (34) and applyultimate boundedness arguments [139]. Note that the right-hand side of (36) is strictly negative for

‖δ‖2 > µc ,‖BTc ω‖2

λ2(L) sinc(ρ)=

λcritical

λ2(L) sinc(ρ).

Pick ε ∈]0, 1[. If ρ ≥ ‖δ‖2 ≥ µc/ε, then the right-hand sideof (36) is upper-bounded by

W (δ) ≤ −(1− ε) · λ2(L) sinc(ρ)W (δ) .

In the following, choose µ such that ρ > µ > µc and let ε =µc/µ ∈ ]0, 1[. By standard ultimate boundedness arguments[139, Theorem 4.18], for ‖δ(0)‖2 ≤ ρ, there is T ≥ 0 suchthat ‖δ(t)‖2 is exponentially decaying for t ∈ [0, T ] and‖δ(t)‖2 ≤ µ for all t ≥ T . For the choice µ = γ withγ ∈ [0, π/2[, the condition µ > µc reduces to

γ sinc(ρ) > λcritical/λ2(L) . (37)

Page 15: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Now, we perform a final analysis of the bound (37). Theleft-hand side of (37) is an increasing function of γ anda decreasing function of ρ. Therefore, there exists some(ρ, γ) in the convex set Λ , (ρ, γ) : ρ ∈ [0, π[ , γ ∈[0, π/2[ , ρ > γ satisfying equation (37) if and only ifthe inequality (37) is true at ρ = γ = π/2, where the left-hand side of (37) achieves its supremum in Λ. The lattercondition is equivalent to inequality (31). Additionally, ifthese two equivalent statements are true, then there is anopen set of points in Λ satisfying (37), which is boundedby the unique curve that satisfies inequality (37) with theequality sign, namely f(ρ, γ) = 0, where f : Λ → R,f(ρ, γ) = γ sinc(ρ)−λcritical/λ2(L). Consequently, for every(ρ, γ) ∈ (ρ, γ) ∈ Λ : f(ρ, γ) > 0, it follows for‖δ(0)‖2 ≤ ρ that there is T ≥ 0 such that ‖δ(t)‖2 ≤ γ for allt ≥ T . The supremum value for ρ is given by ρmax ∈ ]π/2, π]solving the equation f(ρmax, π/2) = 0 and the infimum valueof γ by γmin ∈ [0, π/2[ solving the equation f(γmin, γmin)=0.

This proves statement 1) (where we replaced ρmax by γmax)and shows that there is T ≥ 0 such that ‖BTc θ(t)‖∞ ≤‖BTc θ(t)‖2 ≤ γmin < π/2 for all t ≥ T . Thus, θ(t) ∈∆G(γmin) for t ≥ T , and frequency synchronization can beestablished analogously to the proof of Theorem 4.5.

C. Proof of Theorem 4.7

According to Lemma 4.2, there exists a locally exponen-tially stable synchronization manifold [θ] ∈ ∆G(γ), γ ∈[0, π/2[, if and only if there is an equilibrium θ ∈ ∆G(γ).The equilibrium equations (16) can be rewritten as

ω = L(BT θ)θ , (38)

where L(BT θ) = B diag(aij sinc(θi − θj)i,j∈E)BT isthe Laplacian matrix associated with the graph G(V, E , A)with nonnegative edge weights aij = aij sinc(θi − θj) forθ ∈ ∆G(γ). Since for any weighted Laplacian matrix L, wehave that L · L† = L† · L = In − (1/n)1n×n (follows fromthe singular value decomposition [117]), a multiplication ofequation (38) from the left by BTL(BT θ)† yields

BTL(BT θ)†ω = BT θ . (39)

Note that the left-hand side of equation (39) is a continuous4

function for θ ∈ ∆G(γ). Consider the formal substitutionx = BT θ, the compact and convex set S∞(γ) = x ∈BTRn : ‖x‖∞ ≤ γ, and the continuous map f : S∞(γ)→R given by f(x) = BTL(x)†ω. Then equation (39) readsas the fixed-point equation f(x) = x, and we can invokeBrouwers’s Fixed Point Theorem which states that everycontinuous map from a compact and convex set to itself hasa fixed point, see for instance [143, Section 7, Corollary 8].

Since the analysis of the map f in the ∞-norm is veryhard in the general case, we resort to a 2-norm analysis andrestrict ourselves to the set S2(γ) = x ∈ BTRn : ‖x‖2 ≤γ ⊆ S∞(γ). By Brouwer’s Fixed Point Theorem, there

4 The continuity can be established when re-writing equations (38) and(39) in the quotient space 1⊥n , where L(BT θ) is nonsingular, and usingthe fact that the inverse of a matrix is a continuous function of its elements.

exists a solution x ∈ S2(γ) to the equations x = f(x) if andonly if ‖f(x)‖2 ≤ γ for all x ∈ S2(γ), or equivalently ifand only if

maxx∈S2(γ)

∥∥BTL(x)†ω∥∥2≤ γ . (40)

In the following we show that (32) is a sufficient conditionfor inequality (40).

First, we establish some identities. For a Laplacian matrixL, we obtain L† = V diag(0, 1/λi(L)i=2,...,n)V T , whereλ1(L) = 0 and λi(L) > 0, i ∈ 2, . . . , n, are the eigenval-ues of L and V ∈ Rn×n is an associated orthonormal matrixof eigenvectors. It follows that V diag (0, 1, . . . , 1)V T =In − (1/n)1n×n, and since ω ⊥ 1n, there exists α ∈ R|E|(not necessarily unique), such that ω = Bα. By means ofthese identities, the left-hand side of (40) can be simplifiedand upper-bounded for all x ∈ S2(γ):∥∥BTL(x)†ω

∥∥2

=∥∥BTL(x)†Bα

∥∥2

=∥∥∥∥BTV (x) diag

(0,

1

λ2(L(x)), . . . ,

1

λn(L(x))

)V T (x)Bα

∥∥∥∥2

≤ 1

λ2(L(x))·∥∥BTV (x) diag (0, 1, . . . , 1)V T (x)Bα

∥∥2

= (1/λ2(L(x))) ·∥∥BTω∥∥

2. (41)

Thus, a sufficient condition for inequality (40) to be true canbe derived as follows:

maxx∈S2(γ)

∥∥BTL(x)†ω∥∥2≤∥∥BTω∥∥

2max

x∈S2(γ)

(1/λ2

(L(x)

)))

≤∥∥BTω∥∥

2max

x∈x∈R|E|: ‖x‖∞≤γ

(1/λ2

(L(x)

)))

=∥∥BTω∥∥

2/ (λ2(L) · sinc(γ))

!≤ γ ,

where we used identity (41), we enlarged the domain S2(γ)to x ∈ R|E| : ‖x‖∞ ≤ γ, and we used the fact λ2(L(x)) ≥λ2(L) · sinc(γ) for ‖x‖∞ ≤ γ. In summary, we concludethat there is a locally exponentially stable synchronizationmanifold [θ] ∈ θ ∈ Tn : ‖BT θ‖2≤γ ⊆ ∆G(γ) if

λ2(L) sin(γ) ≥ ‖BTω‖2 . (42)

Since the left-hand side of (42) is a concave function of γ ∈[0, π/2[, there exists an open set of γ ∈ [0, π/2[ satisfyingequation (42) if and only if equation (42) is true with thestrict equality sign at γ∗ = π/2, which corresponds tocondition (32). Additionally, if these two equivalent state-ments are true, then there exists a unique γmin ∈ [0, π/2[that satisfies equation (27) with the equality sign, namelysin(γmin) = ‖BTω‖2/λ2(L). This concludes the proof.

REFERENCES

[1] C. Huygens, Horologium Oscillatorium, Paris, France, 1673.[2] S. H. Strogatz, SYNC: The Emerging Science of Spontaneous Order.

Hyperion, 2003.[3] A. T. Winfree, The Geometry of Biological Time, 2nd ed. Springer,

2001.[4] ——, “Biological rhythms and the behavior of populations of coupled

oscillators,” Journal of Theoretical Biology, vol. 16, no. 1, pp. 15–42,1967.

Page 16: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

[5] Y. Kuramoto, “Self-entrainment of a population of coupled non-linear oscillators,” in Int. Symposium on Mathematical Problems inTheoretical Physics, ser. Lecture Notes in Physics, H. Araki, Ed.Springer, 1975, vol. 39, pp. 420–422.

[6] ——, Chemical Oscillations, Waves, and Turbulence. Springer,1984.

[7] S. H. Strogatz, “From Kuramoto to Crawford: Exploring the onsetof synchronization in populations of coupled oscillators,” Physica D:Nonlinear Phenomena, vol. 143, no. 1, pp. 1–20, 2000.

[8] J. A. Acebron, L. L. Bonilla, C. J. P. Vicente, F. Ritort, and R. Spigler,“The Kuramoto model: A simple paradigm for synchronizationphenomena,” Reviews of Modern Physics, vol. 77, no. 1, pp. 137–185,2005.

[9] D. C. Michaels, E. P. Matyas, and J. Jalife, “Mechanisms of sinoatrialpacemaker synchronization: a new hypothesis,” Circulation Research,vol. 61, no. 5, pp. 704–714, 1987.

[10] C. Liu, D. R. Weaver, S. H. Strogatz, and S. M. Reppert, “Cellu-lar construction of a circadian clock: period determination in thesuprachiasmatic nuclei,” Cell, vol. 91, no. 6, pp. 855–860, 1997.

[11] F. Varela, J. P. Lachaux, E. Rodriguez, and J. Martinerie, “Thebrainweb: Phase synchronization and large-scale integration,” NatureReviews Neuroscience, vol. 2, no. 4, pp. 229–239, 2001.

[12] E. Brown, P. Holmes, and J. Moehlis, “Globally coupled oscillatornetworks,” in Perspectives and Problems in Nonlinear Science: ACelebratory Volume in Honor of Larry Sirovich, E. Kaplan, J. E.Marsden, and K. R. Sreenivasan, Eds. Springer, 2003, pp. 183–215.

[13] S. M. Crook, G. B. Ermentrout, M. C. Vanier, and J. M. Bower, “Therole of axonal delay in the synchronization of networks of coupledcortical oscillators,” Journal of Computational Neuroscience, vol. 4,no. 2, pp. 161–172, 1997.

[14] A. K. Ghosh, B. Chance, and E. K. Pye, “Metabolic coupling andsynchronization of NADH oscillations in yeast cell populations,”Archives of Biochemistry and Biophysics, vol. 145, no. 1, pp. 319–331, 1971.

[15] J. Buck, “Synchronous rhythmic flashing of fireflies. II.” QuarterlyReview of Biology, vol. 63, no. 3, pp. 265–289, 1988.

[16] T. J. Walker, “Acoustic synchrony: two mechanisms in the snowytree cricket,” Science, vol. 166, no. 3907, pp. 891–894, 1969.

[17] N. Kopell and G. B. Ermentrout, “Coupled oscillators and the designof central pattern generators,” Mathematical Biosciences, vol. 90, no.1-2, pp. 87–109, 1988.

[18] N. E. Leonard, T. Shen, B. Nabet, L. Scardovi, I. D. Couzin, and S. A.Levin, “Decision versus compromise for animal groups in motion,”Proceedings of the National Academy of Sciences, vol. 109, no. 1,pp. 227–232, 2012.

[19] D. A. Paley, N. E. Leonard, R. Sepulchre, D. Grunbaum, and J. K.Parrish, “Oscillator models and collective motion,” IEEE ControlSystems Magazine, vol. 27, no. 4, pp. 89–105, 2007.

[20] Z. Neda, E. Ravasz, T. Vicsek, Y. Brechet, and A. L. Barabasi,“Physics of the rhythmic applause,” Physical Review E, vol. 61, no. 6,p. 6987, 2000.

[21] H. Daido, “Quasientrainment and slow relaxation in a population ofoscillators with random and frustrated interactions,” Physical ReviewLetters, vol. 68, no. 7, pp. 1073–1076, 1992.

[22] G. Jongen, J. Anemuller, D. Bolle, A. C. C. Coolen, and C. Perez-Vicente, “Coupled dynamics of fast spins and slow exchange inter-actions in the XY spin glass,” Journal of Physics A: Mathematicaland General, vol. 34, no. 19, pp. 3957–3984, 2001.

[23] J. Pantaleone, “Stability of incoherence in an isotropic gas of os-cillating neutrinos,” Physical Review D, vol. 58, no. 7, p. 073002,1998.

[24] K. Wiesenfeld, P. Colet, and S. H. Strogatz, “Frequency locking inJosephson arrays: Connection with the Kuramoto model,” PhysicalReview E, vol. 57, no. 2, pp. 1563–1569, 1998.

[25] I. Z. Kiss, Y. Zhai, and J. L. Hudson, “Emerging coherence in apopulation of chemical oscillators,” Science, vol. 296, no. 5573, p.1676, 2002.

[26] P. A. Tass, “A model of desynchronizing deep brain stimulation witha demand-controlled coordinated reset of neural subpopulations,”Biological Cybernetics, vol. 89, no. 2, pp. 81–88, 2003.

[27] A. Nabi and J. Moehlis, “Single input optimal control for globallycoupled neuron networks,” Journal of Neural Engineering, vol. 8, p.065008, 2011.

[28] R. Sepulchre, D. A. Paley, and N. E. Leonard, “Stabilization of planar

collective motion: All-to-all communication,” IEEE Transactions onAutomatic Control, vol. 52, no. 5, pp. 811–824, 2007.

[29] ——, “Stabilization of planar collective motion with limited commu-nication,” IEEE Transactions on Automatic Control, vol. 53, no. 3,pp. 706–719, 2008.

[30] D. J. Klein, “Coordinated control and estimation for multi-agentsystems: Theory and practice,” Ph.D. dissertation, University ofWashington, 2008.

[31] D. J. Klein, P. Lee, K. A. Morgansen, and T. Javidi, “Integration ofcommunication and control using discrete time Kuramoto models formultivehicle coordination over broadcast networks,” IEEE Journal onSelected Areas in Communications, vol. 26, no. 4, pp. 695–705, 2008.

[32] M. M. U. Rahman, R. Mudumbai, and S. Dasgupta, “Consensusbased carrier synchronization in a two node network,” in IFAC WorldCongress, Milan, Italy, Aug. 2011, pp. 10 038–10 043.

[33] G. Kozyreff, A. G. Vladimirov, and P. Mandel, “Global coupling withtime delay in an array of semiconductor lasers,” Physical ReviewLetters, vol. 85, no. 18, pp. 3809–3812, 2000.

[34] F. C. Hoppensteadt and E. M. Izhikevich, “Synchronization oflaser oscillators, associative memory, and optical neurocomputing,”Physical Review E, vol. 62, no. 3, pp. 4010–4013, 2000.

[35] R. A. York and R. C. Compton, “Quasi-optical power combiningusing mutually synchronized oscillator arrays,” IEEE Transactionson Microwave Theory and Techniques, vol. 39, no. 6, pp. 1000–1009,2002.

[36] W. C. Lindsey, F. Ghazvinian, W. C. Hagmann, and K. Dessouky,“Network synchronization,” Proceedings of the IEEE, vol. 73, no. 10,pp. 1445–1467, 1985.

[37] O. Simeone, U. Spagnolini, Y. Bar-Ness, and S. H. Strogatz, “Dis-tributed synchronization in wireless networks,” IEEE Signal Process-ing Magazine, vol. 25, no. 5, pp. 81–97, 2008.

[38] Y. W. Hong and A. Scaglione, “A scalable synchronization protocolfor large scale sensor networks and its applications,” IEEE Journalon Selected Areas in Communications, vol. 23, no. 5, pp. 1085–1099,2005.

[39] R. Baldoni, A. Corsaro, L. Querzoni, S. Scipioni, and S. T. Pier-giovanni, “Coupling-based internal clock synchronization for large-scale dynamic distributed systems,” IEEE Transactions on Paralleland Distributed Systems, vol. 21, no. 5, pp. 607–619, 2010.

[40] Y. Wang, F. Nunez, and F. J. Doyle, “Increasing sync rate of pulse-coupled oscillators via phase response function design: theory andapplication to wireless networks,” IEEE Transactions on ControlSystems Technology, 2012, to appear.

[41] E. Mallada and A. Tang, “Distributed clock synchronization: Jointfrequency and phase consensus,” in IEEE Conf. on Decision andControl and European Control Conference, Orlando, FL, USA, Dec.2011, pp. 6742–6747.

[42] S. Barbarossa and G. Scutari, “Decentralized maximum-likelihoodestimation for sensor networks composed of nonlinearly coupled dy-namical systems,” IEEE Transactions on Signal Processing, vol. 55,no. 7, pp. 3456–3470, 2007.

[43] J. W. Simpson-Porco, F. Dorfler, and F. Bullo, “Droop-controlledinverters are Kuramoto oscillators,” in IFAC Workshop on DistributedEstimation and Control in Networked Systems, Santa Barbara, CA,USA, Sep. 2012, to appear.

[44] A. Arenas, A. Dıaz-Guilera, J. Kurths, Y. Moreno, and C. Zhou,“Synchronization in complex networks,” Physics Reports, vol. 469,no. 3, pp. 93–153, 2008.

[45] S. Boccaletti, V. Latora, Y. Moreno, M. Chavez, and D. U. Hwang,“Complex networks: Structure and dynamics,” Physics Reports, vol.424, no. 4-5, pp. 175–308, 2006.

[46] S. H. Strogatz, “Exploring complex networks,” Nature, vol. 410, no.6825, pp. 268–276, 2001.

[47] J. A. K. Suykens and G. V. Osipov, “Introduction to focus issue:Synchronization in complex networks,” Chaos, vol. 18, no. 3, pp.037 101–037 101, 2008.

[48] R. Olfati-Saber, J. A. Fax, and R. M. Murray, “Consensus andcooperation in networked multi-agent systems,” Proceedings of theIEEE, vol. 95, no. 1, pp. 215–233, 2007.

[49] W. Ren, R. W. Beard, and E. M. Atkins, “Information consensus inmultivehicle cooperative control: Collective group behavior throughlocal interaction,” IEEE Control Systems Magazine, vol. 27, no. 2,pp. 71–82, 2007.

[50] F. Bullo, J. Cortes, and S. Martınez, Distributed Control of Robotic

Page 17: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

Networks, ser. Applied Mathematics Series. Princeton UniversityPress, 2009.

[51] L. Moreau, “Stability of multiagent systems with time-dependentcommunication links,” IEEE Transactions on Automatic Control,vol. 50, no. 2, pp. 169–182, 2005.

[52] A. Jadbabaie, N. Motee, and M. Barahona, “On the stability ofthe Kuramoto model of coupled nonlinear oscillators,” in AmericanControl Conference, Boston, MA, USA, Jun. 2004, pp. 4296–4301.

[53] N. Chopra and M. W. Spong, “On exponential synchronization ofKuramoto oscillators,” IEEE Transactions on Automatic Control,vol. 54, no. 2, pp. 353–357, 2009.

[54] Z. Lin, B. Francis, and M. Maggiore, “State agreement forcontinuous-time coupled nonlinear systems,” SIAM Journal on Con-trol and Optimization, vol. 46, no. 1, pp. 288–307, 2007.

[55] A. Sarlette and R. Sepulchre, “Consensus optimization on manifolds,”SIAM Journal on Control and Optimization, vol. 48, no. 1, pp. 56–76,2009.

[56] L. Scardovi, A. Sarlette, and R. Sepulchre, “Synchronization andbalancing on the N -torus,” Systems & Control Letters, vol. 56, no. 5,pp. 335–341, 2007.

[57] R. Olfati-Saber, “Swarms on sphere: A programmable swarm withsynchronous behaviors like oscillator networks,” in IEEE Conf. onDecision and Control, San Diego, CA, USA, 2006, pp. 5060–5066.

[58] G. B. Ermentrout, “An adaptive model for synchrony in the fireflypteroptyx malaccae,” Journal of Mathematical Biology, vol. 29, no. 6,pp. 571–585, 1991.

[59] S. Y. Ha, E. Jeong, and M. J. Kang, “Emergent behaviour ofa generalized Viscek-type flocking model,” Nonlinearity, vol. 23,no. 12, pp. 3139–3156, 2010.

[60] S. Ha, C. Lattanzio, B. Rubino, and M. Slemrod, “Flocking andsynchronization of particle models,” Quarterly Applied Mathematics,vol. 69, pp. 91–103, 2011.

[61] A. R. Bergen and D. J. Hill, “A structure preserving model for powersystem stability analysis,” IEEE Transactions on Power Apparatusand Systems, vol. 100, no. 1, pp. 25–35, 1981.

[62] P. W. Sauer and M. A. Pai, Power System Dynamics and Stability.Prentice Hall, 1998.

[63] H.-D. Chiang, C. C. Chu, and G. Cauley, “Direct stability analysis ofelectric power systems using energy functions: Theory, applications,and perspective,” Proceedings of the IEEE, vol. 83, no. 11, pp. 1497–1529, 1995.

[64] F. Dorfler and F. Bullo, “Synchronization and transient stabilityin power networks and non-uniform Kuramoto oscillators,” SIAMJournal on Control and Optimization, vol. 50, no. 3, pp. 1616–1642,2012.

[65] J. Pantaleone, “Synchronization of metronomes,” American Journalof Physics, vol. 70, p. 992, 2002.

[66] S. H. Strogatz, D. M. Abrams, A. McRobie, B. Eckhardt, andE. Ott, “Theoretical mechanics: Crowd synchrony on the MillenniumBridge,” Nature, vol. 438, no. 7064, pp. 43–44, 2005.

[67] M. Bennett, M. F. Schatz, H. Rockwood, and K. Wiesenfeld, “Huy-gens’s clocks,” Proceedings: Mathematical, Physical and Engineer-ing Sciences, vol. 458, no. 2019, pp. 563–579, 2002.

[68] Y.-P. Choi, S.-Y. Ha, and S.-B. Yun, “Complete synchronization ofKuramoto oscillators with finite inertia,” Physica D, vol. 240, no. 1,pp. 32–44, 2011.

[69] H. A. Tanaka, A. J. Lichtenberg, and S. Oishi, “Self-synchronizationof coupled oscillators with hysteretic responses,” Physica D: Nonlin-ear Phenomena, vol. 100, no. 3-4, pp. 279–300, 1997.

[70] ——, “First order phase transition resulting from finite inertia incoupled oscillator systems,” Physical Review Letters, vol. 78, no. 11,pp. 2104–2107, 1997.

[71] H. Hong, G. S. Jeon, and M. Y. Choi, “Spontaneous phase oscillationinduced by inertia and time delay,” Physical Review E, vol. 65, no. 2,p. 026208, 2002.

[72] H. Hong, M. Y. Choi, J. Yi, and K. S. Soh, “Inertia effects on periodicsynchronization in a system of coupled oscillators,” Physical ReviewE, vol. 59, no. 1, p. 353, 1999.

[73] J. A. Acebron, L. L. Bonilla, and R. Spigler, “Synchronization inpopulations of globally coupled oscillators with inertial effects,”Physical Review E, vol. 62, no. 3, pp. 3437–3454, 2000.

[74] F. Dorfler and F. Bullo, “On the critical coupling for Kuramotooscillators,” SIAM Journal on Applied Dynamical Systems, vol. 10,no. 3, pp. 1070–1099, 2011.

[75] F. De Smet and D. Aeyels, “Partial entrainment in the finiteKuramoto–Sakaguchi model,” Physica D: Nonlinear Phenomena,vol. 234, no. 2, pp. 81–89, 2007.

[76] R. Mirollo and S. H. Strogatz, “The spectrum of the partially lockedstate for the Kuramoto model,” Journal of Nonlinear Science, vol. 17,no. 4, pp. 309–347, 2007.

[77] Y. L. Maistrenko, O. V. Popovych, and P. A. Tass, “Desynchroniza-tion and chaos in the Kuramoto model,” in Dynamics of CoupledMap Lattices and of Related Spatially Extended Systems, ser. LectureNotes in Physics, J.-R. Chazottes and B. Fernandez, Eds. Springer,2005, vol. 671, pp. 285–306.

[78] R. Tonjes, “Pattern formation through synchronization in systems ofnonidentical autonomous oscillators,” Ph.D. dissertation, UniversitatsPotsdam, Germany, 2007.

[79] O. V. Popovych, Y. L. Maistrenko, and P. A. Tass, “Phase chaos incoupled oscillators,” Physical Review E, vol. 71, no. 6, p. 065201,2005.

[80] J. Lunze, “Complete synchronization of Kuramoto oscillators,” Jour-nal of Physics A: Mathematical and Theoretical, vol. 44, p. 425102,2011.

[81] E. Canale and P. Monzon, “Almost global synchronization of sym-metric Kuramoto coupled oscillators,” in Systems Structure andControl. InTech Education and Publishing, 2008, ch. 8, pp. 167–190.

[82] M. Verwoerd and O. Mason, “On computing the critical couplingcoefficient for the Kuramoto model on a complete bipartite graph,”SIAM Journal on Applied Dynamical Systems, vol. 8, no. 1, pp. 417–453, 2009.

[83] S.-Y. Ha, T. Ha, and J.-H. Kim, “On the complete synchronizationof the Kuramoto phase model,” Physica D: Nonlinear Phenomena,vol. 239, no. 17, pp. 1692–1700, 2010.

[84] R. E. Mirollo and S. H. Strogatz, “The spectrum of the lockedstate for the Kuramoto model of coupled oscillators,” Physica D:Nonlinear Phenomena, vol. 205, no. 1-4, pp. 249–266, 2005.

[85] D. Aeyels and J. A. Rogge, “Existence of partial entrainment andstability of phase locking behavior of coupled oscillators,” Progresson Theoretical Physics, vol. 112, no. 6, pp. 921–942, 2004.

[86] M. Verwoerd and O. Mason, “Global phase-locking in finite pop-ulations of phase-coupled oscillators,” SIAM Journal on AppliedDynamical Systems, vol. 7, no. 1, pp. 134–160, 2008.

[87] G. B. Ermentrout, “Synchronization in a pool of mutually coupledoscillators with random frequencies,” Journal of Mathematical Biol-ogy, vol. 22, no. 1, pp. 1–9, 1985.

[88] M. Verwoerd and O. Mason, “A convergence result for the Kuramotomodel with all-to-all coupling,” SIAM Journal on Applied DynamicalSystems, vol. 10, no. 3, pp. 906–920, 2011.

[89] J. C. Bronski, L. DeVille, and M. J. Park, “Fully synchronoussolutions and the synchronization phase transition for the finite nKuramoto model,” Arxiv preprint arXiv:1111.5302, 2011.

[90] J. A. Rogge and D. Aeyels, “Stability of phase locking in a ring ofunidirectionally coupled oscillators,” Journal of Physics A, vol. 37,pp. 11 135–11 148, 2004.

[91] L. DeVille, “Transitions amongst synchronous solutions for thestochastic Kuramoto model,” Nonlinearity, vol. 25, no. 5, pp. 1–20,2011.

[92] U. Munz, A. Papachristodoulou, and F. Allgower, “Consensus reach-ing in multi-agent packet-switched networks with non-linear cou-pling,” International Journal of Control, vol. 82, no. 5, pp. 953–969,2009.

[93] E. Mallada and A. Tang, “Synchronization of phase-coupled os-cillators with arbitrary topology,” in American Control Conference,Baltimore, MD, USA, Jun. 2010, pp. 1777–1782.

[94] L. Scardovi, “Clustering and synchronization in phase models withstate dependent coupling,” in IEEE Conf. on Decision and Control,Atlanta, GA, USA, Dec. 2010, pp. 627–632.

[95] A. Franci, A. Chaillet, and W. Pasillas-Lepine, “Existence androbustness of phase-locking in coupled Kuramoto oscillators undermean-field feedback,” Automatica, vol. 47, no. 6, pp. 1193–1202,2011.

[96] S. Y. Ha and M. Slemrod, “A fast-slow dynamical systems theory forthe Kuramoto type phase model,” Journal of Differential Equations,vol. 251, no. 10, pp. 2685–2695, 2011.

[97] L. Buzna, S. Lozano, and A. Diaz-Guilera, “Synchronization insymmetric bipolar population networks,” Physical Review E, vol. 80,no. 6, p. 66120, 2009.

Page 18: Florian Dorfler Francesco Bullo¨ - arXiv · 2012-09-07 · Exploring Synchronization in Complex Oscillator Networks Florian Dorfler Francesco Bullo¨ Abstract—The emergence of

[98] Y. Moreno and A. F. Pacheco, “Synchronization of Kuramoto oscil-lators in scale-free networks,” Europhysics Letters, vol. 68, p. 603,2004.

[99] A. C. Kalloniatis, “From incoherence to synchronicity in the networkKuramoto model,” Physical Review E, vol. 82, no. 6, p. 066202, 2010.

[100] A. Sarlette, “Geometry and symmetries in coordination control,”Ph.D. dissertation, University of Liege, Belgium, Jan. 2009.

[101] E. A. Canale, P. A. Monzn, and F. Robledo, “The wheels: an infinitefamily of bi-connected planar synchronizing graphs,” in IEEE Conf.Industrial Electronics and Applications, Taichung, Taiwan, Jun. 2010,pp. 2204–2209.

[102] E. A. Canale, P. Monzon, and F. Robledo, “On the complexity ofthe classification of synchronizing graphs,” in Grid and DistributedComputing, Control and Automation, Jeju Island, Korea, Dec. 2010,pp. 186–195.

[103] P. Monzon, “Almost global stability of dynamical systems,” Ph.D.dissertation, Universidad de la Republica, Montevideo, Uruguay, Jul.2006.

[104] J. L. van Hemmen and W. F. Wreszinski, “Lyapunov function forthe Kuramoto model of nonlinearly coupled oscillators,” Journal ofStatistical Physics, vol. 72, no. 1, pp. 145–166, 1993.

[105] S. J. Chung and J. J. Slotine, “On synchronization of coupled Hopf-Kuramoto oscillators with phase delays,” in IEEE Conf. on Decisionand Control, Atlanta, GA, USA, Dec. 2010, pp. 3181–3187.

[106] G. S. Schmidt, U. Munz, and F. Allgower, “Multi-agent speed con-sensus via delayed position feedback with application to Kuramotooscillators,” in European Control Conference, Budapest, Hungary,Aug. 2009, pp. 2464–2469.

[107] E. Canale and P. Monzon, “On the characterization of families ofsynchronizing graphs for Kuramoto coupled oscillators,” in IFACWorkshop on Distributed Estimation and Control in NetworkedSystems, Venice, Italy, Sep. 2009, pp. 42–47.

[108] S. H. Strogatz and R. E. Mirollo, “Phase-locking and critical phenom-ena in lattices of coupled nonlinear oscillators with random intrinsicfrequencies,” Physica D: Nonlinear Phenomena, vol. 31, no. 2, pp.143–168, 1988.

[109] H. Chiba, “A proof of the Kuramoto’s conjecture for a bifurcationstructure of the infinite dimensional Kuramoto model,” Arxiv preprintarXiv:1008.0249, 2010.

[110] E. A. Martens, E. Barreto, S. H. Strogatz, E. Ott, P. So, and T. M.Antonsen, “Exact results for the Kuramoto model with a bimodalfrequency distribution,” Physical Review E, vol. 79, no. 2, p. 26204,2009.

[111] H. Yin, P. G. Mehta, S. P. Meyn, and U. V. Shanbhag, “Synchro-nization of coupled oscillators is a game,” IEEE Transactions onAutomatic Control, vol. 57, no. 4, pp. 920–935, 2012.

[112] D. A. Wiley, S. H. Strogatz, and M. Girvan, “The size of the syncbasin,” Chaos, vol. 16, no. 1, p. 015103, 2006.

[113] Y. Wang and F. J. Doyle, “On influences of global and local cueson the rate of synchronization of oscillator networks,” Automatica,vol. 47, no. 6, pp. 1236–1242, 2011.

[114] F. Dorfler, M. Chertkov, and F. Bullo, “Synchronization in com-plex oscillator networks and smart grids,” Jul. 2012, available athttp://arxiv.org/pdf/1208.0045.

[115] E. W. Justh and P. S. Krishnaprasad, “Equilibria and steering lawsfor planar formations,” Systems & Control Letters, vol. 52, no. 1, pp.25–38, 2004.

[116] S. Sastry and P. Varaiya, “Hierarchical stability and alert state steeringcontrol of interconnected power systems,” IEEE Transactions onCircuits and Systems, vol. 27, no. 11, pp. 1102–1112, 1980.

[117] F. Dorfler and F. Bullo, “Kron reduction of graphs with applicationsto electrical networks,” IEEE Transactions on Circuits and Systems,Nov. 2011, to appear.

[118] V. Fioriti, S. Ruzzante, E. Castorini, E. Marchei, and V. Rosato,“Stability of a distributed generation network using the Kuramotomodels,” in Critical Information Infrastructure Security, ser. LectureNotes in Computer Science. Springer, 2009, pp. 14–23.

[119] G. Filatrella, A. H. Nielsen, and N. F. Pedersen, “Analysis of apower grid using a Kuramoto-like model,” The European PhysicalJournal B, vol. 61, no. 4, pp. 485–491, 2008.

[120] M. Rohden, A. Sorge, M. Timme, and D. Witthaut, “Self-organizedsynchronization in decentralized power grids,” Physical Review Let-ters, vol. 109, no. 6, p. 064101, 2012.

[121] D. Subbarao, R. Uma, B. Saha, and M. V. R. Phanendra, “Self-

organization on a power system,” IEEE Power Engineering Review,vol. 21, no. 12, pp. 59–61, 2001.

[122] D. J. Hill and G. Chen, “Power systems as dynamic networks,” inIEEE Int. Symposium on Circuits and Systems, Kos, Greece, May2006, pp. 722–725.

[123] F. C. Hoppensteadt and E. M. Izhikevich, Weakly connected neuralnetworks. Springer, 1997, vol. 126.

[124] E. M. Izhikevich, Dynamical Systems in Neuroscience: The Geometryof Excitability and Bursting. MIT Press, 2007.

[125] E. M. Izhikevich and Y. Kuramoto, “Weakly coupled oscillators,”Encyclopedia of Mathematical Physics, vol. 5, p. 448, 2006.

[126] G. B. Ermentrout and N. Kopell, “Frequency plateaus in a chainof weakly coupled oscillators, I.” SIAM journal on MathematicalAnalysis, vol. 15, no. 2, pp. 215–237, 1984.

[127] F. F. Wu and S. Kumagai, Limits on Power Injections for PowerFlow Equations to Have Secure Solutions. Electronics ResearchLaboratory, College of Engineering, University of California, 1980.

[128] T. Nishikawa, A. E. Motter, Y. C. Lai, and F. C. Hoppensteadt,“Heterogeneity in oscillator networks: Are smaller worlds easier tosynchronize?” Physical Review Letters, vol. 91, no. 1, p. 14101, 2003.

[129] F. Wu and S. Kumagai, “Steady-state security regions of powersystems,” IEEE Transactions on Circuits and Systems, vol. 29, no. 11,pp. 703–711, 1982.

[130] G. Korniss, M. B. Hastings, K. E. Bassler, M. J. Berryman, B. Kozma,and D. Abbott, “Scaling in small-world resistor networks,” PhysicsLetters A, vol. 350, no. 5-6, pp. 324–330, 2006.

[131] L. M. Pecora and T. L. Carroll, “Master stability functions forsynchronized coupled systems,” Physical Review Letters, vol. 80,no. 10, pp. 2109–2112, 1998.

[132] J. Gomez-Gardenes, Y. Moreno, and A. Arenas, “Paths to synchro-nization on complex networks,” Physical Review Letters, vol. 98,no. 3, p. 34101, 2007.

[133] C. J. Tavora and O. J. M. Smith, “Stability analysis of powersystems,” IEEE Transactions on Power Apparatus and Systems,vol. 91, no. 3, pp. 1138–1144, 1972.

[134] A. Araposthatis, S. Sastry, and P. Varaiya, “Analysis of power-flow equation,” International Journal of Electrical Power & EnergySystems, vol. 3, no. 3, pp. 115–126, 1981.

[135] C. J. Tavora and O. J. M. Smith, “Equilibrium analysis of powersystems,” IEEE Transactions on Power Apparatus and Systems,vol. 91, no. 3, pp. 1131–1137, 1972.

[136] M. Ilic, “Network theoretic conditions for existence and uniquenessof steady state solutions to electric power circuits,” in IEEE Inter-national Symposium on Circuits and Systems, San Diego, CA, USA,May 1992, pp. 2821–2828.

[137] K. S. Chandrashekhar and D. J. Hill, “Cutset stability criterion forpower systems using a structure-preserving model,” InternationalJournal of Electrical Power & Energy Systems, vol. 8, no. 3, pp.146–157, 1986.

[138] R. Sepulchre, A. Sarlette, and P. Rouchon, “Consensus in non-commutative spaces,” in IEEE Conf. on Decision and Control,Atlanta, GA, USA, Dec. 2010, pp. 6596–6601.

[139] H. K. Khalil, Nonlinear Systems, 3rd ed. Prentice Hall, 2002.[140] L. Moreau, “Stability of continuous-time distributed

consensus algorithms,” Feb. 2008, available athttp://arxiv.org/abs/math/0409010v1.

[141] D. Angeli, “A Lyapunov approach to incremental stability properties,”IEEE Transactions on Automatic Control, vol. 47, no. 3, pp. 410–421,2002.

[142] A. Mauroy, “On the dichotomic collective behaviors of large popula-tions of pulse-coupled firing oscillators,” Ph.D. dissertation, Univer-sity of Liege, Belgium, 2011.

[143] E. H. Spanier, Algebraic Topology. Springer, 1994.