This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Generic criterion for explosive synchronization in heterogeneous phase oscillator populations

Can Xu [email protected] Institute of Systems Science and College of Information Science and Engineering, Huaqiao University, Xiamen 361021, China    Xuan Wang Institute of Systems Science and College of Information Science and Engineering, Huaqiao University, Xiamen 361021, China    Per Sebastian Skardal [email protected] Department of Mathematics, Trinity College, Hartford, Connecticut 06106, USA
Abstract

Ordered and disordered behavior in large ensembles of coupled oscillators map to different functional states in a wide range of applications, e.g., active and resting states in the brain and stable and unstable power grid configurations. For this reason, explosive synchronization transitions, which facilitate fast, abrupt changes between these functional states, has recently seen significant interest from researchers. While previous work has identified properties of complex systems that support explosive synchronization, these investigations have been conducted largely on an ad-hoc basis. Here we provide an exact criterion for explosive synchronization in coupled phase oscillator ensembles by investigating the necessary relationship between dynamical and structural disorder. This result provides a critical step towards untangling the intertwined properties of complex systems responsible for abrupt phase transitions and bistability in biological and engineered systems.

Synchronization is a universal phenomenon that emerges in both natural and man-made systems Strogatz2003 ; Pikovsky2003 , and unraveling the underlying mechanism that give rise to emergent self-organizing behavior is of great importance in applications of physical, biological, technological, and social systems Tom2017RMP . The Kuramoto model of couple oscillators Kuramoto1975 , originally proposed in 1975, and its many variants and generalizations, serve as particularly useful paradigms for studying collective behavior. In the classical Kuramoto model the interplay between coupling and heterogeneity of natural frequencies results in a continuous, non-equilibrium phase transition of the Kuramoto order parameter, delineating a partially ordered state from the completely disordered state Strogatz2000PD ; JA2005RMP .

Beyond this continuous phase transition, discontinuous, i.e., explosive, phase transitions towards synchronization have also been observed and have since received ample attention from the nonlinear dynamics and complex networks communities JGG2011prl ; SB2016 ; RDS2019 . The most notable feature of such an abrupt phase transition is the presence of hysteresis that gives rise to a region of bistability between incoherent and coherent states. The implications of explosive synchronization transitions are significant as, most notably, they enable abrupt transitions between ordered and disordered states under relatively small perturbations or parameter changes. In neuroscience and power grid applications, where synchronized vs coherent states characterize active vs resting brain states or stable vs unstable energy configurations, characterizing an exact criterion for the emergence of explosive transitions could be critical for system function EM2015prx ; MA2018PRL ; MR2012PRL . Thus, there is a critical need to understand the impacts of network structures, coupling schemes, and adaptive patterns on the emergence of explosive synchronization in complex systems DP2005PRE ; EA2009PRE ; JGG2011prl ; ZY2014PRL ; JP2013PRL ; XZ2015PRL ; SC2019PRX ; AP2020PRL ; XD2020PRL ; SB2016PR ; RD2019ADV ; CK2021SA ; ZW2019PRE ; XC2019PREE ; CW2021EPJB ; WX2021PRE ; Skardal2014PRE ; Skardal2020CommPhys . Despite substantial attention and observed behavior in many variations and extensions of the Kuramoto model, a mathematically rigorous condition for the origin of explosive synchronization in coupled phase oscillator populations remains lacking.

The aim of this Letter is to fill this critical gap by providing such an exact criterion for explosive synchronization transitions. We focus our attention on phase oscillators with quenched disorder in both the local dynamics and structure, characterized by heterogeneity in natural frequencies and individual coupling strengths, respectively. In particular, we explore the effects that correlations between dynamical and structural disorder have on the resulting collective dynamics. The emergence of an explosive synchronization transition, characterized by a hysteresis loop supporting bistability between synchronized and incoherent states, is established using a self-consistent approach. We uncover the origin of such abrupt transitions by revealing the local bifurcation mechanism of the system near criticality. Specifically, explosive transitions emerge when the concavities of the natural frequency distribution the coupling strength function combine to satisfy a particular inequality. Our results establish a rigorous characterization of the interplay between structure and dynamics that leads to explosive synchronization dynamics.

We consider here a generalized Kuramoto model of the form

θi˙=ωi+KiNj=1Nsin(θjθi),i=1,,N.\dot{\theta_{i}}=\omega_{i}+\frac{K_{i}}{N}\sum_{j=1}^{N}\sin(\theta_{j}-\theta_{i}),\quad i=1,...,N. (1)

Here, θiS1\theta_{i}\in S^{1} denotes the phase of oscillator ii, ωi\omega_{i} is its natural frequency chosen randomly from a prescribed distribution g(ω)g(\omega), N is the size of the system, and KiK_{i} accounts for the heterogeneous interactions between phase oscillators DI2013PRL ; DI2014NC . Thus, the state of the system is described by a flow on the N-torus 𝕋N=(S1)N\mathbb{T}^{N}=(S^{1})^{N}. To establish correlations between the oscillator’s local dynamics and coupling, we set Ki=Kf(ωi)K_{i}=Kf(\omega_{i}) so that the constant K0K\geq 0 controls the overall coupling strength and f(ω)0f(\omega)\geq 0 characterizes correlations between dynamical and structural disorder XU2016PRE . In particular, f(ω)=1f(\omega)=1 recovers the conventional Kuramoto model, while the choice of f(ω)f(\omega) as a step function with mixed signs represents the constrain-conformist model investigated in previous studies HH2011PRL ; HH2016PRE ; HH2016CHAOS . The model defined by Eq. (1) has rich and diverse rhythmic dynamics depending upon g(ω)g(\omega) and f(ω)f(\omega) HB2016PRL ; CX2018PRE ; CX2019NJP ; YJ2021CHAOS ; CX2021PRR .

The aim of this work is to uncover a generic condition for the occurrence of explosive synchronization induced by the correlated disorder described above. For this purpose we restrict our attention to the case where g(ω)g(\omega) and f(ω)f(\omega) are even functions, i.e., symmetric about their mean, which we may set to ω=0\omega=0 without loss of generality by entering an appropriate reference frame, and twice continuously differentiable at ω=0\omega=0, but make no further restrictions on g(ω)g(\omega) and f(ω)f(\omega). To quantify the overall degree of synchronization we use the complex order parameter Z(t)=R(t)eiΘ(t)=N1j=1Neiθj(t)Z(t)=R(t)e^{i\Theta(t)}=N^{-1}\sum_{j=1}^{N}e^{i\theta_{j}(t)}, for which the amplitude R(t)[0,1]R(t)\in[0,1] measures the coherence of the system, and the argument Θ(t)S1\Theta(t)\in S^{1} locates the average phase of the population.

We first illustrate an exact result giving a criterion for the emergence of explosive synchronization. To this aim, we focus on finding the equilibrium states assuming that the system Eq. (1) evolves to a steady state in tt\rightarrow\infty. More precisely, we assume that R(t)R(t) approaches a constant value and Θ(t)\Theta(t) processes at a constant angular velocity. In fact, by setting the mean frequency to ω=0\omega=0, we may assume that this angular velocity is zero and then by shifting initial conditions we may set Θ(t)=0\Theta(t)=0. Therefore, the governing equation Eq. (1) can be rewritten using the mean-field,

θ˙=ωqf(ω)sinθ,\dot{\theta}=\omega-qf(\omega)\sin\theta, (2)

where q=KRq=KR and the index ii is dropped in the continuum limit, NN\rightarrow\infty.

Next, at steady-state the phase oscillators split into two distinct groups depending on their frequencies. Specifically, those satisfying |ω|<qf(ω)|\omega|<qf(\omega) become phase-locked, i.e., entrained by the mean-field, and ultimately converge to a state satisfying sinθω=ω/[qf(ω)]\sin\theta_{\omega}=\omega/[{qf(\omega)}] and cosθω=1sin2θω\cos\theta_{\omega}=\sqrt{1-\sin^{2}\theta_{\omega}}. On the other hand, those satisftying |ω|>qf(ω)|\omega|>qf(\omega) never come to rest and remain drifting on S1S^{1} for all time. The order parameter is expressed as Z=eiθl+eiθdZ=\langle e^{i\theta}\rangle_{l}+\langle e^{i\theta}\rangle_{d}, where l\langle{\cdot}\rangle_{l} and d\langle{\cdot}\rangle_{d} represent the ensemble average over the locked and drifting populations, respectively. It can be shown that, due to the symmetry of gg and ff, eiθd==0\langle e^{i\theta}\rangle_{d}=\langle=0 and sinθl=0\sin{\theta}\rangle_{l}=0, which implies that the drifting populations have no contribution to the order parameter, and the locked ones contribute to its real part only RM2007NS . Thus, the self-consistent equation describing the steady state of the system is Z=R=cosθlZ=R=\langle\cos{\theta}\rangle_{l}, which can be reformulated in a simple form given by

1K=H(q)=1q|ω|<qf(ω)1ω2q2f2(ω)g(ω)𝑑ω.\frac{1}{K}=H(q)=\frac{1}{q}\int_{|\omega|<qf(\omega)}\sqrt{1-\frac{\omega^{2}}{q^{2}f^{2}(\omega)}}g(\omega)d\omega. (3)

Setting F(ω)=ω/f(ω)F(\omega)=\omega/{f(\omega)}, and defining x=F(ω)/qx={F(\omega)}/{q}, the function H(q)H(q) may be simplified to

H(q)=111x2G(qx)𝑑x,H(q)=\int_{-1}^{1}\sqrt{1-x^{2}}G(qx)dx, (4)

where G(qx)G(qx) is denoted the generalized distribution (virtual distribution) GJ2020PRE given by

G(qx)=g[F1(qx)]f[F1(qx)]1qxf[F1(qx)],G(qx)=\frac{g[F^{-1}(qx)]f[F^{-1}(qx)]}{1-qxf^{{}^{\prime}}[F^{-1}(qx)]}, (5)

F1(qx)F^{-1}(qx) is the inverse function determined by qx=F(ω)qx=F(\omega), and the apostrophe represents the derivative. The self-consistent equations Eq. (3)–(5) characterize the degree of synchronization at steady-state. Since the incoherent state corresponds to q=0q=0, G(0)=g(0)f(0)G(0)=g(0)f(0), the critical coupling strength KcK_{c} characterizing the onset of synchronization via the instability of the incoherent state may be identified by taking the limit q0+q\rightarrow 0^{+}, giving

Kc=2πg(0)f(0).K_{c}=\frac{2}{\pi g(0)f(0)}. (6)

Remarkably, Eq. (6) is a straightforward generalization of the classical result obtained by Kuramoto Kuramoto1975 , now with f(0)f(0) modifying the onset of synchronization.

Next, to identify an explosive synchronization transition, which is marked by a subcritical bifurcation at K=KcK=K_{c} (contrasting a continuous transition, which is marked by a supercritical one), we let K=Kc+δKK=K_{c}+\delta K and q=0+δqq=0+\delta q, with δK,δq1\delta K,\delta q\ll 1. Expanding both sides of Eq. (3) yields

1KcδKKc2=H(0)+H(0)δq+12H′′(0)δq2.\frac{1}{K_{c}}-\frac{\delta K}{K_{c}^{2}}=H(0)+H^{\prime}(0)\delta q+\frac{1}{2}H^{\prime\prime}(0)\delta q^{2}. (7)

The first terms on either side of Eq. (7) balance due to Eq. (6), and the second term on the right hand side of Eq. (7) vanishes due to the symmetry of G(qx)G(qx). Using the relation δq2Kc2δR2\delta q^{2}\approx K^{2}_{c}\delta R^{2}, we obtain that at criticality the order parameter is given by XC2020PRE ,

δR=2δKH′′(0)Kc4.\delta R=\sqrt{\frac{-2\delta K}{H^{\prime\prime}(0)K^{4}_{c}}}. (8)

We conclude from Eq. (8) that H′′(0)>0H^{\prime\prime}(0)>0 is the desired condition for subcriticality, i.e., explosive synchronization, where the phase transition from incoherence to synchronization is abrupt. On the other hand, when H′′(0)<0H^{\prime\prime}(0)<0 the bifurcation supercritical, yielding a continuous phase transition. In fact, we have H′′(0)=11x21x2G′′(0)𝑑x=π8G′′(0)H^{\prime\prime}(0)=\int_{-1}^{1}x^{2}\sqrt{1-x^{2}}G^{\prime\prime}(0)dx=\frac{\pi}{8}G^{\prime\prime}(0), so the value G′′(0)G^{\prime\prime}(0) suffices to dictate the nature of the transition at K=KcK=K_{c}.

To obtain a more explicit condition for explosive synchronization, we expand g(ω)g(\omega) and f(ω)f(\omega) at the meanω=0\omega=0, i.e., g(ω)=g(0)+g′′(0)2ω2+𝒪(ω4)g(\omega)=g(0)+\frac{g^{\prime\prime}(0)}{2}\omega^{2}+\mathcal{O}(\omega^{4}) and f(ω)=f(0)+f′′(0)2ω2+𝒪(ω4)f(\omega)=f(0)+\frac{f^{\prime\prime}(0)}{2}\omega^{2}+\mathcal{O}(\omega^{4}). Without loss of generality, f(0)f(0) can be set to one by rescaling the time tt. To leading order we then have F1(qx)=(112f′′(0)q2x2)/[f′′(0)qx]F^{-1}(qx)=({1-\sqrt{1-2f^{\prime\prime}(0)q^{2}x^{2}}})/[{f^{\prime\prime}(0)qx}], with 0<qx10<qx\ll 1. The generalized distribution G(qx)G(qx) near qx=0qx=0 is then given by, to leading order,

G(qx)=ω[g(0)+21g′′(0)]qx[1qxf′′(0)ω]|ω=F1(qx).G(qx)=\frac{\omega[g(0)+2^{-1}g^{\prime\prime}(0)]}{qx[1-qxf^{\prime\prime}(0)\omega]}\Big{|}_{\omega=F^{-1}(qx)}. (9)

Using F1(qx)F^{-1}(qx), we eventually obtain G′′(0)=g′′(0)+3g(0)f′′(0)G^{\prime\prime}(0)=g^{\prime\prime}(0)+3g(0)f^{\prime\prime}(0), yielding a universal condition for the onset of explosive synchronization given by

g′′(0)+3g(0)f′′(0)>0.g^{\prime\prime}(0)+3g(0)f^{\prime\prime}(0)>0. (10)
Refer to caption
Figure 1: Continuous vs explosive synchronization transitions. The order parameter RR vs. coupling strength KK for g(ω)=γ/[π(ω2+γ2)]g(\omega)=\gamma/[\pi(\omega^{2}+\gamma^{2})] with γ=0.5\gamma=0.5 and f(ω)=2eβω2f(\omega)=2-e^{-\beta\omega^{2}}, giving βc=4/3\beta_{c}=4/3. Panels (a)–(c) correspond to β<βc\beta<\beta_{c}, yielding continuous transitions, while panels (d)–(f) corresponds to β>βc\beta>\beta_{c} exhibiting explosive transitions. Red and blue circles indicate results from numerical simulations with N=105N=10^{5} oscillators, adiabatically increasing and decreasing KK, respectively. The solid (stable) and dashed (unstable) lines represent theoretical predictions, respectively.

Once the inequality in Eq. (10) is satisfied, the phase transition towards synchronization is explosive. For example, if g(ω)=γ/[π(ω2+γ2)]g(\omega)={\gamma}/[{\pi}({\omega^{2}+\gamma^{2}})] and f(ω)=2eβω2f(\omega)=2-e^{-\beta\omega^{2}} the critical parameter for the onset of explosive synchronization is βc=1/(3γ2)\beta_{c}=1/({3\gamma^{2}}), above which the transition is abrupt. Using γ=1/2\gamma=1/2, yielding a critical value of βc=4/3\beta_{c}=4/3 we illustrate our main result in Fig. 1(a)–(f), plotting for β=0\beta=0, 0.50.5, 11, 22, 88, and 1414 the synchronization transitions RR vs KK. Results from direct simulations with N=105N=10^{5} oscillators are plotted in red and blue circles, indicating simulations where KK are adiabatically increased and decreased, respectively, and theoretical curves are plotted in black. Note that all simulations with β<βc\beta<\beta_{c} (top row) yields continuous transitions, and those with β>βc\beta>\beta_{c} (bottom row) yield explosive transitions. More generally, for the typical case of a unimodal frequency distribution with g′′(0)<0g^{\prime\prime}(0)<0, it follows that the transition to synchronization is explosive only for sufficiently positive f′′(0)f^{\prime\prime}(0), specifically, f′′(0)>g′′(0)/[3g(0)]f^{\prime\prime}(0)>-g^{\prime\prime}(0)/[3g(0)]. Note that, given the conditions on f(ω)f(\omega), this requires that there exists some open interval containing ω=0\omega=0 where f(ω)f(\omega) is both concave up as well as monotonically increasing as |ω||\omega| increases. Thus, in some interval near the mean frequency, explosive synchronization requires a sufficiently positive correlation between the natural frequencies and local coupling strengths.

For the remainder of this letter we present a rigorous bifurcation analysis of the onset of synchronization that occurs at K=KcK=K_{c}, putting our main result on a rigorous mathematical foundation. We begin by noting that in the continuum limit the governing equation Eq. (1) is equivalent to the following continuity equation

ρt+(ρv)θ=0\frac{\partial\rho}{\partial t}+\frac{\partial(\rho v)}{\partial\theta}=0 (11)

with velocity field v(θ,ω,t)=ω+Kf(ω)Im(Zeiθ)v(\theta,\omega,t)=\omega+{Kf(\omega)}\mathrm{Im}(Ze^{-i\theta}) and ρ(θ,ω,t)\rho(\theta,\omega,t) representing the real, 2π2\pi-periodic distribution describing the density of oscillators with phase θ\theta and natural frequency ω\omega at time tt whose nthn^{\text{th}} Fourier coefficient of ρ\rho is defined by αn=02πeinθρ(θ,ω,t)𝑑θ\alpha_{n}=\int_{0}^{2\pi}e^{in\theta}\rho(\theta,\omega,t)d\theta for with α0=1\alpha_{0}=1 and αn(ω,t)=α¯n(ω,t){\alpha}_{-n}(\omega,t)={\bar{\alpha}}_{n}(\omega,t), where the over-bar denotes complex conjugate.

The Ott-Antonsen ansatz restricts the Fourier coefficients to a special form OTT2008CHAOS ; OTT2009CHAOS defined by αn(ω,t)=αn(ω,t){\alpha}_{n}(\omega,t)={\alpha}^{n}(\omega,t) for n1n\geq 1, which is an attracting, invariant manifold of Eq. (11). On this manifold, the low dimensional dynamics of the system evolve according to

dαdt=iωα+Kf(ω)2(ZZ¯α2),\frac{d\alpha}{dt}=i\omega\alpha+\frac{Kf(\omega)}{2}(Z-\bar{Z}\alpha^{2}), (12)

which is closed with the order parameter Z(t)Z(t) defined by the integral

Z(t)=+α(ω,t)g(ω)𝑑ω=𝒢α(ω,t).Z(t)=\int_{-\infty}^{+\infty}\alpha(\omega,t)g(\omega)d\omega=\mathcal{G}\alpha(\omega,t). (13)

(The integral operate 𝒢\mathcal{G} is introduced to ease notation.) The steady state solutions of the low dimensional dynamics in Eq. (12) are expressed as a simple profile OE2013PD ,

αs(ω)={iω+q2f2(ω)ω2qf(ω),|ω|qf(ω),iωsgn(ω)ω2q2f2(ω)qf(ω),|ω|>qf(ω),\alpha_{s}(\omega)=\begin{cases}\frac{i\omega+\sqrt{q^{2}f^{2}(\omega)-\omega^{2}}}{qf(\omega)},&|\omega|\leq qf(\omega),\\ i\frac{\omega-\mathrm{sgn}(\omega)\sqrt{\omega^{2}-q^{2}f^{2}(\omega)}}{qf(\omega)},&|\omega|>qf(\omega),\end{cases} (14)

where sgn(ω)\mathrm{sgn}(\omega) is the sign function with respect to ω\omega, and the first and second terms corresponds to the phase-locked and drifting populations, respectively.

To analyze the stability of the steady solutions α(ω)\alpha(\omega) we consider small perturbations away from αs(ω)\alpha_{s}(\omega), namely, α(ω,t)=αs(ω)+εv(ω,t)\alpha(\omega,t)=\alpha_{s}(\omega)+\varepsilon v(\omega,t), where 0<ε10<\varepsilon\ll 1 and v(ω,t)v(\omega,t) is the perturbation vector. Then, the order parameter may be written Z(t)=R+ε𝒢v(ω,t)Z(t)=R+\varepsilon\mathcal{G}v(\omega,t) and the linearized dynamics around αs(ω)\alpha_{s}(\omega) is

dvdt=ηωv+K2(aω𝒢v+bω𝒢v¯),\frac{dv}{dt}=\eta_{\omega}v+\frac{K}{2}(a_{\omega}\mathcal{G}v+b_{\omega}\mathcal{G}\bar{v}), (15)

where ηω=iωqf(ω)αs(ω)\eta_{\omega}=i\omega-qf(\omega)\alpha_{s}(\omega), and the coefficients aω=f(ω)a_{\omega}=f(\omega) , bω=f(ω)αs2(ω)b_{\omega}=-f(\omega)\alpha_{s}^{2}(\omega). By introducing 𝐕=(v,v¯)T\mathbf{V}=(v,\bar{v})^{T} we transform Eq. (15) to

d𝐕dt=𝐌𝐕+K2Qg^𝐕,\frac{d\mathbf{V}}{dt}=\mathbf{M}\mathbf{V}+\frac{K}{2}Q\hat{{g}}\mathbf{V}, (16)

where 𝐌=(ηω00η¯ω)\mathbf{M}=\left(\begin{array}[]{cccc}\eta_{\omega}&0\\ 0&\bar{\eta}_{\omega}\\ \end{array}\right), and 𝐐=(aωbωb¯ωa¯ω)\mathbf{Q}=\left(\begin{array}[]{cccc}a_{\omega}&b_{\omega}\\ \bar{b}_{\omega}&\bar{a}_{\omega}\\ \end{array}\right). Using d𝐕/dt=λ𝐕d\mathbf{V}/dt=\lambda\mathbf{V} we obtain

𝐕=K2(λ𝐈𝐌)1𝐐𝒢𝐕,\mathbf{V}=\frac{K}{2}(\lambda\mathbf{I}-\mathbf{M})^{-1}\mathbf{Q}\mathcal{G}\mathbf{V}, (17)

where 𝐈\mathbf{I} is the unit matrix. Applying the integral operate 𝒢\mathcal{G} to both sides of Eq. (17) leads to the following two dimensional linear equations

(𝐈K𝐉)𝒢𝐕=0,(\mathbf{I}-K\mathbf{J})\mathcal{G}\mathbf{V}=0, (18)

with

𝐉=12𝒢[(λ𝐈𝐌)1𝐐].\mathbf{J}=\frac{1}{2}\mathcal{G}[(\lambda\mathbf{I}-\mathbf{M})^{-1}\mathbf{Q}]. (19)

A nontrivial solution of Eq. (18) requires that det(𝐈K𝐉)=0\det(\mathbf{I}-K\mathbf{J})=0, from which we arrive at the pair of eigenvalue equations

1K=hc(λ)\displaystyle\frac{1}{K}=h_{c}(\lambda) =12𝒢(1αs2)f(ω)ληω,\displaystyle=\frac{1}{2}\mathcal{G}\frac{(1-\alpha_{s}^{2})f(\omega)}{\lambda-\eta_{\omega}}, (20)
1K=hs(λ)\displaystyle\frac{1}{K}=h_{s}(\lambda) =12𝒢(1+αs2)f(ω)ληω,\displaystyle=\frac{1}{2}\mathcal{G}\frac{(1+\alpha_{s}^{2})f(\omega)}{\lambda-\eta_{\omega}}, (21)

where λ\lambda\in\mathbb{R} and ληω\lambda\neq\eta_{\omega}.

Recall that, for the incoherent state αs(ω)=0\alpha_{s}(\omega)=0, ηω=iω\eta_{\omega}=i\omega, we have

hc(λ)=hs(λ)=12+λf(ω)g(ω)λ2+ω2𝑑ω.h_{c}(\lambda)=h_{s}(\lambda)=\frac{1}{2}\int_{-\infty}^{+\infty}\frac{\lambda f(\omega)g(\omega)}{\lambda^{2}+\omega^{2}}d\omega. (22)

In the limit λ0+\lambda\rightarrow 0^{+}, the function λ/(λ2+ω2)πδ(0)\lambda/(\lambda^{2}+\omega^{2})\rightarrow\pi\delta(0), yielding, KKc=2/[πg(0)f(0)]K\rightarrow K_{c}=2/[\pi g(0)f(0)], which coincides with onset of synchronization obtained previously. Thus, the phase transition towards synchronization takes place at the critical coupling strength KcK_{c} at which point the incoherent state loses stability.

As for the partially synchronized state with αs(ω)0\alpha_{s}(\omega)\neq 0, the sign of λ\lambda controls their stability properties. It can be shown that hs(0)=q1𝒢α0(ω)h_{s}(0)=q^{-1}\mathcal{G}\alpha_{0}(\omega), which is exactly the self-consistent equation Eq. (3). This implies that λ=0\lambda=0 is always a trivial root of Eq. (21) originating form the rotational symmetry of Eq. (1). However, for λ0\lambda\neq 0, we define Δ(λ)=hc(λ)K1\Delta(\lambda)=h_{c}(\lambda)-K^{-1}, which satisfies

Δ(0)=qH(q).\Delta(0)=qH^{\prime}(q). (23)

To prove Eq. (23) we note that

qH(q)=q[1q𝒢αs(ω)]=𝒢[dαs(ω)dq]1K.qH^{\prime}(q)=q[\frac{1}{q}\mathcal{G}\alpha_{s}(\omega)]^{\prime}=\mathcal{G}[\frac{d\alpha_{s}(\omega)}{dq}]-\frac{1}{K}. (24)

Following Eq. (14), we have that, for |ω|<qf(ω)|\omega|<qf(\omega),

dαsdq=isin2θωqcosθω,\frac{d\alpha_{s}}{dq}=i\frac{\sin^{2}{\theta}_{\omega}}{q\cos{\theta}_{\omega}}, (25)

and for |ω|>qf(ω)|\omega|>qf(\omega),

dαsdq=i[f(ω)sgn(ω)ω2q2f2(ω)ωsgn(ω)ω2q2f2(ω)q2f(ω)].\frac{d\alpha_{s}}{dq}=i[\frac{f(\omega)\text{sgn}(\omega)}{\sqrt{\omega^{2}-q^{2}f^{2}(\omega)}}-\frac{\omega-\text{sgn}(\omega)\sqrt{\omega^{2}-q^{2}f^{2}(\omega)}}{q^{2}f(\omega)}]. (26)

Putting all this together, we obtain

hc(0)=𝒢(dαsdq),h_{c}(0)=\mathcal{G}\left(\frac{d\alpha_{s}}{dq}\right), (27)

which completes the proof of Eq. (23). In Ref. CX2021PREE it was shown that Δ(0)>0\Delta(0)>0 implies instability. By contrast, Δ(0)<0\Delta(0)<0 indicates that the corresponding stationary solution described by H(q)H(q) remains asymptotically stable. According to Eq. (7), H(q)H′′(0)qH^{\prime}(q)\sim H^{\prime\prime}(0)q for 0<q10<q\ll 1. Hence, for H′′(0)>0H^{\prime\prime}(0)>0 the associated branch of the bifurcating solutions is unstable, and in the limit q0+q\rightarrow 0^{+} this unstable branch collides with the stable incoherent state, thereby creating a subcritical bifurcation at KcK_{c}, i.e., an explosive transition to synchronization.

Take together, we have fully characterized the dynamics of a generalized Kuramoto model, in which the interactions are deterministically correlated with the given oscillator’s intrinsic frequency. We have revealed a universal condition that leads to the explosive synchronization via hysteresis and bistability. Furthermore, we have demonstrated the mathematical basis for such an abrupt transition by developing the local bifurcation theory of the incoherent state at onset. Our study could find applicability in better understanding the nonequilibrium phase transitions and related critical phenomenon in complex systems.

ACKNOWLEDGEMENTS

CX and XW acknowledge support from the National Natural Science Foundation of China (Grants No. 11905068) and the Scientific Research Funds of Huaqiao University (Grant No. ZQN-810). PSS acknowledges support from the National Science Foundation (Grant No. MCB-2126177). We thank the support from Higher-Order Network Reading Group supported by the Save 2050 Programme jointly sponsored by Swarma Club and X-Order.

References

  • (1) S. H. Strogatz, Sync: The Emerging Science of Spontaneous Order (Hypernion, New York, 2003).
  • (2) A. Pikovsky, M. Rosenblum, and J. Kurths, Synchronization: a universal concept in nonlinear sciences (Cambridge University Press, 2003).
  • (3) T. Stankovski, T.o Pereira, P. V. E. McClintock, and A. Stefanovska, Coupling functions: Universal insights into dynamical interaction mechanisms, Rev. Mod. Phys. 89, 045001 (2017).
  • (4) Y. Kuramoto, in International Symposium on Mathematical Problems in Theoretical Physics, edited by H. Araki, Lecture Notes in Physics, No. 30 (Springer, New York, 1975), p. 420.
  • (5) S. H. Strogatz, From Kuramoto to Crawford: Exploring the onset of synchronization in populations of coupled oscillators, Physica D 143, 1 (2000).
  • (6) J. A. Acebron, L. L. Bonilla, C. J. Perez Vicente, F. Ritort, and R. Spigler, The Kuramoto model: A simple paradigm for synchronization phenomena, Rev. Mod. Phys. 77, 137 (2005).
  • (7) J. Gómez-Gardeñes, S. Gómez, A. Arenas, and Y. Moreno, Explosive synchronization transitions in scale-free networks, Phys. Rev. Lett. 106, 128701 (2011).
  • (8) S. Boccaletti, J. A. Almendral, S. Guan, I.Leyva, Z. Liu, I.Sendiña-Nadal, Z. Wang, and Y. Zou, Explosive transitions in complex networks’ structure and dynamics: Percolation and synchronization, Phys. Rep. 660, 1 (2016).
  • (9) R. D’Souza, J. Nagler, J. Gómez-Gardeñes and A. Arenas, Explosive phenomena in complex networks, Adv. Phys. 68(3), 123 (2019).
  • (10) E. Montbrió, D. Pazó, and A. Roxin, Macroscopic Description for Networks of Spiking Neurons, Phys. Rev. X 5, 021028 (2015).
  • (11) M. di Volo and A. Torcini, Transition from Asynchronous to Oscillatory Dynamics in Balanced Spiking Networks with Instantaneous Synapses, Phys. Rev. Lett. 121, 128301 (2018).
  • (12) M. Rohden, A. Sorge, M. Timme, and D. Witthaut, Self-Organized Synchronization in Decentralized Power Grids, Phys. Rev. Lett. 109, 064101 (2012).
  • (13) D. Pazó, Thermodynamic limit of the first-order phase transition in the Kuramoto model, Phys. Rev. E 72, 046211 (2005).
  • (14) E. A. Martens, E. Barreto, S. H. Strogatz, E. Ott, P. So, and T. M. Antonsen, Exact results for the Kuramoto model with a bimodal frequency distribution, Phys. Rev. E 79, 026204 (2009).
  • (15) Y. Zou, T. Pereira, M. Small, Z. Liu, and Jürgen Kurths, Basin of Attraction Determines Hysteresis in Explosive Synchronization, Phys. Rev. Lett. 112, 114102 (2014).
  • (16) P. Ji, T. K. DM. Peron, P. J. Menck, F. A. Rodrigues, J. and Kurths, Cluster explosive synchronization in complex networks, Phys. Rev. Lett. 110, 218701 (2013).
  • (17) X. Zhang, S. Boccaletti, S. Guan, and Z. Liu, Explosive Synchronization in Adaptive and Multilayer Networks, Phys. Rev. Lett. 114, 038701 (2015).
  • (18) S. Chandra, M. Girvan, and E. Ott, Continuous versus Discontinuous Transitions in the D-Dimensional Generalized Kuramoto Model: Odd D is Different, Phys. Rev. X 9, 011002 (2019).
  • (19) A. P. Millán, J. J. Torres, and G. Bianconi, Explosive Higher Order Kuramoto Dynamics on Simplicial Complexes, Phys. Rev. Lett. 124, 218301 (2020).
  • (20) X. Dai, X. Li, H. Guo, D. Jia, M. Perc, P. Manshour, Z. Wang, and S. Boccaletti, Discontinuous Transitions and Rhythmic States in the D-Dimensional Kuramoto Model Induced by a Positive Feedback with the Global Order Parameter, Phys. Rev. Lett. 125, 194101 (2020).
  • (21) S. Boccaletti, J. A. Almendral, S. Guan, I. Leyva, Z. Liu, I. Sendia-Nadal, Z. Wang, and Y. Zou, Explosive transitions in complex networks structure and dynamics: Percolation and synchronization, Phys. Rep. 660, 1 (2016).
  • (22) R. D’ Souza, J. Gómez-Gardeñes, J. Nagler, and A. Arenas, Explosive phenomena in complex networks, Adv. Phys. 68, 123 (2019).
  • (23) C. Kuehn and C. Bick, A universal route to explosive phenomena, Sci. Adv. 7, eabe3824 (2021).
  • (24) W. Zou, M. Zhan, and J. Kurths, Phase transition to synchronization in generalized Kuramoto model with low-pass filter, Phys. Rev. E 100, 012209 (2019).
  • (25) C. Xu, J. Gao, S. Boccaletti, Z. Zheng, and S. Guan, Synchronization in starlike networks of phase oscillators, Phys. Rev. E 100, 012212 (2019).
  • (26) W. Chen, S. Wang, Y. Lan, et al. Explosive synchronization caused by optimizing synchrony of coupled phase oscillators on complex networks. Eur. Phys. J. B 94, 205 (2021).
  • (27) X. Wang, Z. Zheng, and C. Xu, Collective dynamics of phase oscillator populations with three-body interactions, Phys. Rev. E 104, 054208 (2021).
  • (28) P. S. Skardal and A. Arenas, Disorder induces explosive synchronization. Phys. Rev. E 89, 062811 (2014).
  • (29) P. S. Skardal and A. Arenas, Higher-order interactions in complex networks of phase oscillators promote abrupt synchronization switching, Comm. Phys. 3, 218 (2020).
  • (30) D. Iatsenko, S. Petkoski, P. V. E. McClintock, and A. Stefanovska, Stationary and Traveling Wave States of the Kuramoto Model with an Arbitrary Distribution of Frequencies and Coupling Strengths, Phys. Rev. Lett. 110, 064101 (2013).
  • (31) D. Iatsenko, P. V. E. McClintock, and A. Stefanovska, Glassy states and super-relaxation in populations of coupled phase oscillators, Nat. Commun. 5, 4118 (2014).
  • (32) C. Xu, J. Gao, H. Xiang, W. Jia, S. Guan, and Z. Zheng, Dynamics of phase oscillators with generalized frequency-weighted coupling, Phys. Rev. E 94, 062204 (2016).
  • (33) H. Hong and S. H. Strogatz, Kuramoto Model of Coupled Oscillators with Positive and Negative Coupling Parameters: An Example of Conformist and Contrarian Oscillators, Phys. Rev. Lett. 106, 054102 (2011).
  • (34) H. Hong, K. P. O’Keeffe, and S. H. Strogatz, Phase coherence induced by correlated disorder, Phys. Rev. E 93, 022219 (2016).
  • (35) Hong H, O’Keeffe KP, Strogatz SH. , Correlated disorder in the Kuramoto model: Effects on phase coherence, finite-size scaling, and dynamic fluctuations. Chaos 26(10), 103105 (2016).
  • (36) H. Bi, X. Hu, S. Boccaletti, X. Wang, Y. Zou, Z. Liu, and S. Guan, Coexistence of Quantized, Time Dependent, Clusters in Globally Coupled Oscillators, Phys. Rev. Lett. 117, 204101 (2016).
  • (37) C. Xu, S. Boccaletti, S. Guan, and Z. Zheng, Origin of Bellerophon states in globally coupled phase oscillators, Phys. Rev. E 98, 050202(R) (2018).
  • (38) C. Xu, S. Boccaletti, Z. Zheng, and S. Guan, Universal phase transitions to synchronization in Kuramoto-like models with heterogeneous coupling, New J. Phys. 21, 113018 (2019).
  • (39) L. Lei, W. Han, J. Yang, Kuramoto model with correlation between coupling strength and natural frequency, Chaos, Solitons & Fractals 144, 110734 (2021).
  • (40) C. Xu, X. Tang, H. Lü, K. Alfaro-Bittner, S. Boccaletti, M. Perc, and S. Guan, Collective dynamics of heterogeneously and nonlinearly coupled phase oscillators, Phys. Rev. Research 3, 043004 (2021).
  • (41) R. Mirollo and S. H. Strogatz, The spectrum of the partially locked state for the Kuramoto model, J. Nonlinear Sci. 17, 309 (2007).
  • (42) J. Gao, and K. Efstathiou, Reduction of oscillator dynamics on complex networks to dynamics on complete graphs through virtual frequencies, Phys. Rev. E 101, 022302 (2020).
  • (43) C. Xu, X. Wang, and P. S. Skardal, Universal scaling and phase transitions of coupled phase oscillator populations, Phys. Rev. E 102, 042310 (2020).
  • (44) E. Ott and T. M. Antonsen, Low dimensional behavior of large systems of globally coupled oscillators, Chaos 18, 037113 (2008).
  • (45) E. Ott and T. M. Antonsen, Long time evolution of phase oscillator systems, Chaos 19, 023117 (2009).
  • (46) O. E. Omel’chenko and M. Wolfrum, Bifurcations in the Sakaguchi-Kuramoto model, Physica D (Amsterdam, Netherlands) 263, 74 (2013).
  • (47) S. H. Strogatz and R. E. Mirollo, Stability of incoherence in a population of coupled oscillators, J. Stat. Phys. 63, 613 (1991).
  • (48) C. Xu, X. Wang, Z. Zheng, and Z. Cai, Stability and bifurcation of collective dynamics in phase oscillator populations with general coupling, Phys. Rev. E 103, 032307 (2021).