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

Stochastic Kuramoto oscillators with inertia and higher-order interactions

Priyanka Rajwani phd2201151017@iiti.ac.in    Sarika Jalan sarika@iiti.ac.in: Corresponding Author Complex Systems Lab, Department of Physics, Indian Institute of Technology Indore, Khandwa Road, Simrol, Indore-453552, India
Abstract

Impact of noise in coupled oscillators with pairwise interactions has been extensively explored. Here, we study stochastic second-order coupled Kuramoto oscillators with higher-order interactions, and show that as noise strength increases the critical points associated with synchronization transitions shift toward higher coupling values. By employing the perturbation analysis, we obtain an expression for the forward critical point as a function of inertia and noise strength. Further, for overdamped systems we show that as noise strength increases, the first-order transition switches to second-order even for higher-order couplings. We include a discussion on nature of critical points obtained through Ott-Antonsen ansatz.

Introduction:

Synchronization is a fundamental phenomenon observed across physics, biology, chemistry, and engineering; from the rhythmic flashing of fireflies, the coordinated firing of neurons in the brain, and the stability of power grids [1]. The Kuramoto model provides a pivotal analytical tool for studying synchronization [2, 3], and demonstrating how a system of interacting oscillators with diverse natural frequencies begins moving in unison as the interaction strength varies. This model is particularly valued for its simplicity and analytical tractability, which not only make theoretical analyses feasible but also enhance its utility in practical applications such as power system [4] and biological systems [5]. Additionally, research has explored how the adaptation function and phase lag parameter affect the synchronized state [6, 7, 8, 9]. Introducing noise to the Kuramoto model incorporates stochasticity, which reflects the intrinsic fluctuations found in real-world systems. Thus, this approach enables more accurate and realistic simulations, effectively mimicking real-world environments [10]. Campa and Gupta considered the case of heterogeneous noise and analyzed the impact of noise strength on dynamical evolution of coupled Kuramoto oscillators [12]. There exist few other studies investigating the impact of noise strength on synchronization profile in networks of phase oscillators [13, 10], and on the model used to study stability of large human connectome graph [11].

Furthermore, incorporation of inertia in the coupled Kuramoto system provides an application for modeling power grids [14, 15, 16]. An inclusion of inertia term in the Kuramoto model has been shown to lead first-order phase transitions characterized by abrupt changes in system dynamics in response to a small change in coupling strength [17]. Managing external perturbations or noise is crucial to prevent systemic failures in many complex systems [18]. Acebrón et. al. studied the influence of noise on critical transition points at which systems dynamics exhibit significant changes [19, 20]. Gupta et. al. outlined a detailed phase space diagram for coupled Kuramoto oscillators with pairwise interactions [21, 22]. Additionally, Cao et. al. described the effects of noise on cluster explosive synchronization in second-order Kuramoto oscillators on networks [23].

All these results were limited to Kuramoto model (with or without inertia) with pairwise interactions. Recent studies have emphasized importance of higher-order interactions in modeling real-world complex systems [24, 25, 26]. Incorporating higher-order interactions into the Kuramoto model has been shown to lead abrupt (de)synchronization transition [27, 28] and tiered synchronization [29, 30] in contrast to second-order transitions observed in pairwise interactions. In absence of noise, the second-order Kuramoto model with higher-order interactions revealed the presence of prolonged hysteresis [31]. Furthermore, studies on the second-order Kuramoto model with higher-order interactions, incorporating phase lag and coupling strengths ranging from negative to positive, have demonstrated the emergence of synchronization and frequency chimera states [32].

This Letter considers the second-order Kuramoto model with higher-order interactions and Gaussian white noise. The study investigates noise-induced transitions in the model. Using Fokker-Plank equation, we first derive distribution function for frequency and phases for identical oscillators, and observe that the frequency order parameter is modulated by inertia and remains unaffected by coupling strength of both pairwise and higher-order interactions. Second, we show that an increase in noise strength shifts the forward (backward) critical point associated with an abrupt jump from an incoherent to coherent state (vice versa) toward higher coupling values. By employing a perturbation analysis, we obtain the expression for critical coupling strength in the forward direction which shows a dependence on inertia and noise strength. Additionally, in an overdamped system, we analytically predict all (un)stable states using the Ott-Antonsen approach. Also, in presence of higher-order interactions, note a shift from first-order to second-order phase transitions as noise strength increases.

Model:

We consider a stochastic Kuramoto model with 22-simplex interactions and inertia. The equation of motion of NN globally coupled oscillators is given as,

mθi¨=θi˙+Ωi+K1Nj=1Nsin(θjθi)+K2N2j=1Nk=1Nsin(2θjθkθi)+ξi(t),\begin{split}m\ddot{\theta_{i}}=&-\dot{\theta_{i}}+{\Omega_{i}}+\frac{K_{1}}{N}\sum_{j=1}^{N}\sin(\theta_{j}-\theta_{i})\\ &+\frac{K_{2}}{N^{2}}\sum_{j=1}^{N}\sum_{k=1}^{N}\sin(2\theta_{j}-\theta_{k}-\theta_{i})+\xi_{i}(t),\end{split} (1)

where θi\theta_{i} and Ωi\Omega_{i} indicate the phase and intrinsic frequency, respectively, of ithi^{th} Kuramoto oscillator. The parameters K1K_{1} and K2K_{2} denote the coupling strength for pairwise and 22-simplex interactions, respectively, and mm is the inertia term. The noise term ξi\xi_{i} is defined as white Gaussian noise, with ξi(t)=0\langle\xi_{i}(t)\rangle=0 and ξi(t)ξj(s)=2Dδijδ(ts)\langle\xi_{i}(t)\xi_{j}(s)\rangle=2D\delta_{ij}\delta(t-s), where DD represents the noise strength. The order parameter is defined as:

rpeιψp=1Nj=1Neιpθj,{r_{p}e^{\iota\psi_{p}}=\frac{1}{N}\sum_{j=1}^{N}e^{\iota p\theta_{j}}}, (2)
Refer to caption
Refer to caption Refer to caption
Figure 1: (Color online) (a) s(t)s(t) vs tt for different values of K1=5K_{1}=5 (orange, ), 10 (violet, ), 15 (red, ), at K2=5K_{2}=5. b) s(t)s(t) vs tt for K1=10K_{1}=10 with different values of K2=0K_{2}=0 (violet, ) and 10 (orange, ). (c) s(t)s(t) vs tt for different values of mm with D=1D=1 and K1=K2=5K_{1}=K_{2}=5. (d) Distribution of frequency η(ω)\eta(\omega) plotted numerically (symbols) and analytically (solid lines) using Eq. Model: and Eq. 9, respectively.

Here, for p=1p=1, 0r110\leq r_{1}\leq 1 quantifies the magnitude of global synchronization, and ψ1\psi_{1} is the mean phase of all oscillators. r1=0r_{1}=0 implies a state in which all oscillators move incoherently around the circle, while r1=1r_{1}=1 indicates global synchronization. Moreover, the frequency order parameter is defined as follows,

seιϕ=1Nj=1Neιωj,{se^{\iota\phi}=\frac{1}{N}\sum_{j=1}^{N}e^{\iota\omega_{j}}}, (3)

where, 0s10\leq s\leq 1 represents the magnitude of frequency synchronization. We express Eq. 1 in the mean-field form by using the order parameter (Eq. 2) such that each oscillator interacts with the mean phase of all the other oscillators,

mθi¨=θi˙+Ωi+K1r1sin(ψ1θi)+K2r2r1sin(ψ2ψ1θi)+ξi(t).\begin{split}m\ddot{\theta_{i}}=-\dot{\theta_{i}}+&{\Omega_{i}}+{K_{1}r_{1}}\sin(\psi_{1}-\theta_{i})\\ &+{K_{2}r_{2}r_{1}}\sin(\psi_{2}-\psi_{1}-\theta_{i})+\xi_{i}(t).\end{split} (4)

Further, expressing Eq. 4 as a system of simultaneous first-order differential equations we get,

θi˙\displaystyle\dot{\theta_{i}} =ωi,\displaystyle=\omega_{i},
ωi˙\displaystyle\dot{\omega_{i}} =1m[ωi+Ωi+K1r1sin(ψ1θi)\displaystyle=\frac{1}{m}[-\omega_{i}+\Omega_{i}+K_{1}r_{1}\sin(\psi_{1}-\theta_{i})
+K2r2r1sin(ψ2ψ1θi)+ξi(t)].\displaystyle\quad+K_{2}r_{2}r_{1}\sin(\psi_{2}-\psi_{1}-\theta_{i})+\xi_{i}(t)]. (5)

Probability density function using Fokker-Plank equation:

We employ the Fokker-Planck equation (FPE) to derive the probability density function ρ(θ,ω,Ω,t)\rho(\theta,\omega,\Omega,t) for the oscillators,

Refer to caption
Figure 2: (Color online) r1r_{1} vs K1K_{1} at (m=0.05m=0.05) illustrating shifts in KcK_{c} with increasing the noise strength DD. (a) K2=0K_{2}=0 and, (b) K2=5K_{2}=5. Results obtained numerically by simulating Eq. Model: with Lorentzian frequency distribution (Ω0=0&Δ=0.5\Omega_{0}=0\,\,\&\,\,\Delta=0.5). Open and filled symbols correspond to forward and backward numerical simulation predictions.
ρt=Dm22ρω21mω[(ω+Ω+K1r1sin(ψ1θ)K2r2r1sin(ψ2ψ1θ))ρ]ωρθ.\begin{split}\frac{\partial\rho}{\partial t}=\frac{D}{m^{2}}\frac{\partial^{2}\rho}{\partial\omega^{2}}-\frac{1}{m}\frac{\partial}{\partial\omega}[(-\omega+\Omega+K_{1}r_{1}\sin(\psi_{1}-\theta)\\ K_{2}r_{2}r_{1}\sin(\psi_{2}-\psi_{1}-\theta))\rho]-\omega\frac{\partial\rho}{\partial\theta}.\end{split} (6)

In the stationary state ρt=0\frac{\partial\rho}{\partial t}=0, let us first consider that all oscillators are identical g(Ω)=δ(Ω)g(\Omega)=\delta(\Omega). Further, the density distribution is 2π2\pi periodic in θ\theta and decays as ω±\omega\rightarrow\pm\infty. Following the normalization condition 02πρ(θ,ω,Ω,0)𝑑ω𝑑θ=1\int_{0}^{2\pi}\int_{-\infty}^{\infty}\rho(\theta,\omega,\Omega,0)d\omega d\theta=1. We look for the solution of the density function in the form of ρ(θ,ω)=η(ω)χ(θ)\rho(\theta,\omega)=\eta(\omega)\chi(\theta), ultimately obtaining the expression for the distribution function of frequencies and phases. Moreover, simulation of Eq. Model: shows that the frequency order parameter is unaffected by K1K_{1} and K2K_{2} (Figs. 1). Hence, solving Eq. 6 in the following manner,

Dm2d2ηdω2+ωmdηdω+ηm=0.\frac{D}{m^{2}}\frac{d^{2}\eta}{d\omega^{2}}+\frac{\omega}{m}\frac{d\eta}{d\omega}+\frac{\eta}{m}=0. (7)
ωηdχdθ1m[K1r1sin(ψ1θ)+K2r2r1sin(ψ2ψ1θ)]χdηdω=0.\begin{split}-\omega\eta\frac{d\chi}{d\theta}-\frac{1}{m}&\Bigl{[}K_{1}r_{1}\sin(\psi_{1}-\theta)+K_{2}r_{2}r_{1}\\ &\sin(\psi_{2}-\psi_{1}-\theta)\Bigr{]}\chi\frac{d\eta}{d\omega}=0.\end{split} (8)
Refer to caption
Figure 3: (Color online) r1r_{1} as function of K1K_{1} for m=1m=1. (a) K2=0K_{2}=0, and (b) K2=5K_{2}=5 for different values of noise strength DD. Here, D=0D=0 (violet, square), D=0.5D=0.5 (red, circle), D=1D=1 (maroon, triangle). Results obtained by simulating Eq. Model: with Lorentzian frequency distribution (Ω0=0,Δ=1\Omega_{0}=0,\,\,\Delta=1). Open and filled symbols correspond to forward and backward numerical simulation predictions, respectively.

By solving these differential equations, and applying boundary and normalization conditions, the distribution functions of frequencies and phases can be represented as follows,

η(ω)=m2πDemω22D.\eta(\omega)=\sqrt{\frac{m}{2\pi D}}e^{\frac{-m\omega^{2}}{2D}}. (9)
χ(θ)=eK1r1cos(ψ1θ)+K2r2r1cos(ψ2ψ1θ)02πeK1r1cos(ψ1θ)+K2r2r1cos(ψ2ψ1θ)𝑑θ.\chi(\theta)=\frac{e^{K_{1}r_{1}\cos(\psi_{1}-\theta)+K_{2}r_{2}r_{1}\cos(\psi_{2}-\psi_{1}-\theta)}}{\int_{0}^{2\pi}e^{K_{1}r_{1}\cos(\psi_{1}-\theta)+K_{2}r_{2}r_{1}\cos(\psi_{2}-\psi_{1}-\theta)}d\theta}. (10)

Fig. 1(d) illustrates the frequency distribution function derived analytically using Eq. 9 (solid lines), which match with the numerically obtained results from Eq. Model: (symbols).

Perturbation analysis of incoherent state:

In the incoherent state, phases are randomly distributed around the unit radius circle implying r1=r2=0r_{1}=r_{2}=0. Using Eq. 6 to obtain the density function of the incoherent state given as,

ρ0(ω,Ω)=12πm2πDem(ωΩ)22D.\rho_{0}(\omega,\Omega)=\frac{1}{2\pi}\sqrt{\frac{m}{2\pi D}}e^{\frac{-m(\omega-\Omega)^{2}}{2D}}. (11)

Giving a small perturbation (ϵ<<1\epsilon<<1) to the incoherent state, we get

ρ(θ,ω,Ω,t)=ρ0(ω,Ω)+ϵα(θ,ω,Ω,t)+O(ϵ2).\rho(\theta,\omega,\Omega,t)=\rho_{0}(\omega,\Omega)+\epsilon\alpha(\theta,\omega,\Omega,t)+O(\epsilon^{2}). (12)

The order parameter in the continuum limit can be written as,

r1eιψ1=02πeιϕρ(ϕ,ω,Ω,t)g(Ω)𝑑Ω𝑑ω𝑑ϕ.r2eιψ2=02πe2ιϕρ(ϕ,ω,Ω,t)g(Ω)𝑑Ω𝑑ω𝑑ϕ.\begin{split}r_{1}e^{\iota\psi_{1}}&=\int_{0}^{2\pi}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{\iota\phi}\rho(\phi,\omega,\Omega,t)g(\Omega)d\Omega d\omega d\phi.\\ r_{2}e^{\iota\psi_{2}}&=\int_{0}^{2\pi}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{2\iota\phi}\rho(\phi,\omega,\Omega,t)g(\Omega)d\Omega d\omega d\phi.\end{split} (13)

Further, incorporating Eqs.( 12 and  13) into Eq. 6, and equating like terms in ϵ\epsilon we obtain

αt+ωαθDm22αω21mω((ωΩ)α)=K1mρ0ω02πsin(ϕθ)α(ϕ,ω,Ω,t)g(Ω)𝑑Ω𝑑ω𝑑ϕ.\frac{\partial\alpha}{\partial t}+\omega\frac{\partial\alpha}{\partial\theta}-\frac{D}{m^{2}}\frac{\partial^{2}\alpha}{\partial\omega^{2}}-\frac{1}{m}\frac{\partial}{\partial\omega}\left((\omega-\Omega)\alpha\right)=-\frac{K_{1}}{m}\frac{\partial\rho_{0}}{\partial\omega}\int_{0}^{2\pi}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\sin{(\phi-\theta)}\alpha(\phi,\omega,\Omega,t)g(\Omega)d\Omega d\omega d\phi. (14)

This equation is exactly the same as in Ref. [20] (Eq. 9), and therefore the subsequent derivations are also the same as in Ref. [20]. For Lorentzian frequency distribution g(Ω)=Δπ[(ΩΩ0)2+Δ2]g(\Omega)=\frac{\Delta}{\pi[(\Omega-\Omega_{0})^{2}+\Delta^{2}]} with mean Ω0=0\Omega_{0}=0 and standard deviation Δ\Delta, we get Kc=2Δ(mΔ+1)+2(2+3mΔ)2+mΔD+O(D2)K_{c}=2\Delta(m\Delta+1)+\frac{2(2+3m\Delta)}{2+m\Delta}D+O(D^{2}) for the forward transition point where oscillators jump from an incoherent to a synchronized state, applicable when noise strength is small (D1D\ll 1). For a negligible inertia mm, Kc=2(D+Δ)K_{c}=2(D+\Delta) is similar to the case of the Kuramoto Model with Gaussian white noise [33]. We find that KcK_{c} is independent of K2K_{2}.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 4: (Color online) Analysis of critical points and hysteresis width as functions of noise strength DD. (a) Forward KcK_{c} and backward K1bK_{1b} critical points for K2=5K_{2}=5 at different inertia values: m=0.5m=0.5 (golden, square) and m=1m=1 (violet, circle). (b) Transition points for m=1m=1 with varying K2K_{2} values: K2=5K_{2}=5 (maroon, circle) and K2=8K_{2}=8 (green, square). (c) For m=1m=1: Δ=0.5\Delta=0.5 (sky blue, square) and Δ=1\Delta=1 (orange, circle). (d) Hysteresis width WW vs DD. Here, (m=0.5,K2=5,Δ=1)(m=0.5,\,\,K_{2}=5,\,\,\Delta=1) (golden, square), (m=1,K2=5,Δ=0.5)(m=1,\,\,K_{2}=5,\,\,\Delta=0.5) (magenta, triangle), (m=1,K2=5,Δ=1)(m=1,\,\,K_{2}=5,\,\,\Delta=1) (violet, circle), and (m=1,K2=8,Δ=1)(m=1,\,\,K_{2}=8,\,\,\Delta=1) (orange, diamond). These results obtain numerically for N=3×104N=3\times 10^{4} using Eq. Model:. For (a), (b), and (c), open and filled symbols represent the forward and backward directions, respectively. Solid lines represent the fitted curve lines.

Numerical Results:

Numerical simulations are performed using the Euler method with a time step dt=0.05dt=0.05, for Eq. 1 by converting it to mean-field equation and then simultaneous first-order differential equations Eq. Model:. Results are obtained for N=20,000N=20,000 and r1r_{1} averaged over 8×1048\times 10^{4} iterations after removing an initial transient by changing K1K_{1} in the forward and backward directions.

Fig. 2 depicts the phase transition for low mm value indicating that critical transition point KcK_{c} shifts towards higher positive coupling value with an increase in DD. Fig. 2(a) represents the second-order phase transition in absence of higher-order interactions (K2=0K_{2}=0). We consider the Lorentzian frequency distribution with standard deviation Δ=0.5\Delta=0.5, so Kc=2D+1K_{c}=2D+1 correctly determines the transition point, for (D<1D<1). Furthermore, Fig. 2(b) for K2=5K_{2}=5 demonstrates that phase transition changes from first-order to second-order as we increase DD, despite the presence of higher-order interactions.

For m=1m=1, Fig. 3 (a, b) depicts the first-order phase transition. With an increase in DD, KcK_{c} shifts towards higher value. Also, the backward transition point K1bK_{1b} at which the oscillators jump from synchronized to incoherent states shifts towards higher value. Further, Fig. 3 (b) depicts that KcK_{c} is not affected by the presence of 22-simplex interactions term, as yielded by the perturbation analysis of the incoherent state that expression of KcK_{c} comes out to be independent of K2K_{2}. Since K2K_{2} impacts K1bK_{1b}, thereby leading to an increase in the hysteresis width compared to the system with only pairwise interactions.

Fig. 4 analyzes forward KcK_{c} and backward K1bK_{1b} transition points as a function of noise strength DD, and examines the impact of changes in mm, K2K_{2}, Δ\Delta. Fig. 4(a) depicts that an increase in inertia mm shifts KcK_{c} towards higher values, while K1bK_{1b} remains unaffected thereby increasing the hysteresis width. Fig. 4(b) demonstrates that increasing K2K_{2} leads to a shift in K1bK_{1b} towards lower values with KcK_{c} remaining the same. This also results in an increased hysteresis width. Fig. 4(c) illustrates that an increase in Δ\Delta value shifts both KcK_{c} and K1bK_{1b} towards higher values. As reflected by Fig. 4(d) hysteresis width W=KcK1bW=K_{c}-K_{1b} decreases as DD increases.

Overdamped system:

In overdamped systems, where the inertia mm is negligible, the Ott-Antonsen approach [34] facilitates the dimensional reduction for intrinsic frequency drawn from the Lorentzian distribution as:

r1˙=(D+Δ)r1+K1(r1r13)2+K2(r13r15)2.\dot{r_{1}}=-(D+\Delta)r_{1}+\frac{K_{1}(r_{1}-r_{1}^{3})}{2}+\frac{K_{2}(r_{1}^{3}-r_{1}^{5})}{2}. (15)
Refer to caption
Figure 5: (Color online) r1r_{1} as a function of K1K_{1} for different values of DD and K2K_{2}. (a) K2=0K_{2}=0 and (b) K2=5K_{2}=5 depict shifts in KcK_{c}. These results are obtained numerically (Eq. 4) for m=0m=0, with Lorentzian frequency distribution (Ω0=0,Δ=1\Omega_{0}=0,\,\,\Delta=1). Forward and backward directions are represented by filled and open symbols, respectively. Analytical results (Eq. 15) are illustrated with solid (stable) and dashed lines (unstable).

Eq. 15 is derived by equating the coefficient of eιθe^{\iota\theta} in the context of pairwise interactions only (as discussed in [35]). We extend the analysis to include 22-simplex interactions. The forward critical transition point Kc=2(D+Δ)K_{c}=2(D+\Delta), is a threshold where oscillators jump from an incoherent to a synchronized state, and for K1>KcK_{1}>K_{c}, r1=0r_{1}=0 solution becomes unstable. By equating the coefficients of eιnθe^{\iota n\theta}, where nn\in\mathbb{N}, we get r1˙=(nD+Δ)r1+K1(r1r13)2+K2(r13r15)2\dot{r_{1}}=-(nD+\Delta)r_{1}+\frac{K_{1}(r_{1}-r_{1}^{3})}{2}+\frac{K_{2}(r_{1}^{3}-r_{1}^{5})}{2}. As we can see that for the noisy case nn arrives explicitly in r1˙\dot{r_{1}} expression, hence Ott-Antonsen approach does not result in the dimensional reduction of Eq. 1 for overdamped system. However, by substituting n=1n=1 in r1˙\dot{r_{1}}, the analytical predictions closely matches with those achieved through numerical simulations (Fig. 5).

Fig. 5 (a) K2=0K_{2}=0 shows second-order phase transition, and (b) K2=5K_{2}=5 manifests first-order phase transition. Increasing DD results in decrease in the hysteresis width, finally yielding second-order phase transition.

Conclusion:

This study investigates noise-induced phase transitions in the Kuramoto model with inertia and 22-simplex interactions. First, using Fokker-Plank equation, we determine the distribution function of frequency and phase for identical oscillators, demonstrating that the frequency order parameter while depends on inertia mm, remains unaffected by coupling strengths K1K_{1} and K2K_{2}. Our analysis further reveals that noise strength is a critical factor governing both forward and backward transition points. Using the perturbation analysis for Lorentzian frequency distribution, we show that the forward transition point is independent of the 22-simplex interaction term and rather depends on inertia and noise strength. By employing the Ott-Antonsen approach we derive approximate analytical predictions for the overdamped system. A significant finding is that as noise strength increases in the overdamped system, the phase transition shifts from the first-order to the second-order even in the presence of higher-order interactions. These results underscore the interplay between noise, inertia, and higher-order interactions in synchronized systems.

We hope that understanding these dynamical behaviours may be useful for studying stability of power grid systems. By incorporating noise and higher-order interactions into the Kuramoto model, we can more accurately predict critical transition points which are essential for preventing systemic failures. This improved modeling approach can aid in designing of more robust power grid systems, capable of maintaining stability under perturbations and noise conditions. Furthermore, few direct possible extensions of this work are to have more realistic noise models [36] and incorporation of multiplicative noise [37].

Acknowledgement:

SJ and PR acknowledge Govt of India SERB Power grant SPF/2021/000136 and PMRF grant PMRF/2023/2103358, respectively.

References

  • [1] G. V. Osipov, J. Kurths, & C. Zhou, Synchronization in Oscillatory Networks, Springer Science & Business Media, (2007).
  • [2] Y. Kuramoto, In International Symposium on Mathematical Problems in Theoretical Physics, Springer Berlin Heidelberg, pp. 420-422, (1975).
  • [3] S. H. Strogatz, Physica D: Nonlinear Phenomena, 143, 1 (2000).
  • [4] Y. Guo, D. Zhang, Z. Li, Q. Wang, & D. Yu, International Journal of Electrical Power & Energy Systems, 129, 106804 (2021).
  • [5] C. Bick, M. Goodfellow, C. R. Laing, & E. A. Martens, The Journal of Mathematical Neuroscience, 10 (1), 9 (2020).
  • [6] P. Khanra, P. Kundu, P. Pal, P. Ji, & C. Hens, Chaos: An Interdisciplinary Journal of Nonlinear Science, 30 (3) (2020).
  • [7] O. E. Omel’Chenko & M. Wolfrum, Physical Review Letters, 109 (16), 164101 (2012).
  • [8] M. Manoranjani, V. R. Saiprasad, R. Gopal, D. V. Senthilkumar, & V. K. Chandrasekar, Physical Review E, 108 (4), 044307 (2023).
  • [9] D. Biswas & S. Gupta, Physical Review E, 109 (2), 024221 (2024).
  • [10] J. Hindes, I. B. Schwartz, & M. Tyloo, Chaos: An Interdisciplinary Journal of Nonlinear Science, 33 (11) (2023).
  • [11] G. Ódor, J. Kelling, & G. Deco, Neurocomputing, 461, 696-704 (2021).
  • [12] A. Campa & S. Gupta, Physical Review E, 108 (6), 064124 (2023).
  • [13] R. K. Esfahani, F. Shahbazi, & K. A. Samani, Physical Review E—Statistical, Nonlinear, and Soft Matter Physics, 86 (3), 036204 (2012).
  • [14] G. Filatrella, A. H. Nielsen, & N. F. Pedersen, The European Physical Journal B, 61, 485-491 (2008).
  • [15] H. A. Tanaka, A. J. Lichtenberg, & S. Oishi, Physical Review Letters, 78, 2104-2107 (1997).
  • [16] M. Rohden, A. Sorge, M. Timme, & D. Witthaut, Physical Review Letters, 109 (6), 064101 (2012).
  • [17] J. Gao & K. Efstathiou, Physical Review E, 98 (4), 042201 (2018).
  • [18] L. Tumash, S. Olmi, & E. Schöll, Europhysics Letters, 123 (2), 20001 (2018); S. Olmi, A. Navas, S. Boccaletti, & A. Torcini, Physical Review E, 90 (4), 042905 (2014).
  • [19] J. A. Acebrón & R. Spigler, Physical Review Letters, 81 (11), 2229 (1998).
  • [20] J. A. Acebrón, L. L. Bonilla, & R. Spigler, Physical Review E, 62 (3), 3437 (2000).
  • [21] S. Gupta, A. Campa, & S. Ruffo, Physical Review E, 89 (2), 022123 (2014).
  • [22] M. Komarov, S. Gupta, & A. Pikovsky, Europhysics Letters, 106 (4), 40003 (2014).
  • [23] L. Cao, C. Tian, Z. Wang, X. Zhang, & Z. Liu, Physical Review E, 97 (2), 022220 (2018).
  • [24] F. Battiston, G. Cencetti, I. Iacopini, V. Latora, M. Lucas, A. Patania, et. al., Physics Reports, 874, 1-92 (2020).
  • [25] S. Boccaletti, P. De Lellis, C. I. del Genio, K. Alfaro-Bittner, R. Criado, S. Jalan, and M. Romance, Physics Reports, 1018, 1-64 (2023).
  • [26] Z. Gao, D. Ghosh, H. Harrington, J. Restrepo, and D. Taylor, Chaos: An Interdisciplinary Journal of Nonlinear Science, 33 (2023).
  • [27] P. S. Skardal & A. Arenas, Communications Physics, 3 (1), 218 (2020).
  • [28] S. Jalan and A. Suman, Physical Review E, 106, 044304 (2022).
  • [29] P. S. Skardal & C. Xu, Chaos: An Interdisciplinary Journal of Nonlinear Science, 32 (5) (2022).
  • [30] P. Rajwani, A. Suman, and S. Jalan, Chaos: An Interdisciplinary Journal of Nonlinear Science, 33 (2023).
  • [31] N. G. Sabhahit, A. S. Khurd, & S. Jalan, Physical Review E, 109 (2), 024212 (2024).
  • [32] P. Jaros, S. Ghosh, D. Dudkowski, S. K. Dana, & T. Kapitaniak, Physical Review E, 108 (2), 024215 (2023).
  • [33] S. H. Strogatz & R. E. Mirollo, Journal of Statistical Physics, 63, 613-635 (1991).
  • [34] E. Ott & T. M. Antonsen, Chaos: An Interdisciplinary Journal of Nonlinear Science, 18, (2008).
  • [35] A. Sinha & A. Ghosh, Europhysics Letters, 141 (5), 53001 (2023).
  • [36] B. C. Bag, K. G. Petrosyan, & C. K. Hu, Physical Review E, 76 (5), 056210 (2007).
  • [37] P. D. Pinto, A. L. Penna, & F. A. Oliveira, Europhysics Letters, 117 (5), 50009 (2017).