Random heterogeneity outperforms design in network synchronization
Edited by Michelle Girvan, University of Maryland, College Park, MD, and accepted by Editorial Board Member Herbert Levine April 19, 2021 (received for review November 30, 2020)
Significance
Synchronization among interacting entities is a process that underlies the function of numerous systems, including circadian clocks and laser arrays. It is generally believed that homogeneity among the entities is beneficial for synchronization. This work shows theoretically, numerically, and experimentally that the opposite is not only possible but also common in systems with interaction delays. In such systems, heterogeneity among the entities is shown to promote synchronization, even when the heterogeneity is completely random. This finding advances our understanding of the interplay between order and disorder in the collective behavior of complex systems. We suggest that the phenomenon can be observed for diverse coupling schemes and has implications for real-world systems, where heterogeneity and delays are common and often unavoidable.
Abstract
A widely held assumption on network dynamics is that similar components are more likely to exhibit similar behavior than dissimilar ones and that generic differences among them are necessarily detrimental to synchronization. Here, we show that this assumption does not generally hold in oscillator networks when communication delays are present. We demonstrate, in particular, that random parameter heterogeneity among oscillators can consistently rescue the system from losing synchrony. This finding is supported by electrochemical-oscillator experiments performed on a multielectrode array network. Remarkably, at intermediate levels of heterogeneity, random mismatches are more effective in promoting synchronization than parameter assignments specifically designed to facilitate identical synchronization. Our results suggest that, rather than being eliminated or ignored, intrinsic disorder in technological and biological systems can be harnessed to help maintain coherence required for function.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
Heterogeneity among interacting components is usually seen as detrimental to the emergence of uniform dynamics in networks, including consensus (1, 2) and synchronization (3, 4). For networks of coupled oscillators, the assumption has been that global synchronization would be hindered by parameter mismatches among oscillators. This assumption, which is observed to hold for Kuramoto models (3), remains undersubstantiated for more general classes of oscillator networks (5–7), especially those studied using the master stability function formalism and its variants (8–11). In the relatively few theoretical studies that have explicitly considered oscillator heterogeneity beyond the context of Kuramoto models, the focus has been on small parameter mismatches and the persistence of synchronization among nearly identical oscillators (12–16). These results all conform to the perception that disorder, in the form of random oscillator heterogeneity, is undesirable for synchronization.
Yet, a few exceptions to this perception exist in the literature. In particular, it has been shown that disorder can sometimes enhance synchronization and/or spatiotemporal order in arrays of driven dissipative pendulums (17–19). For example, for coupled oscillators in a chaotic regime, heterogeneity can suppress chaos, giving rise to more regular patterns (17). In these studies, and numerous subsequent ones (20–22), disorder does not stabilize the system around an original synchronization orbit, but instead generates new behavior that is qualitatively different. However, it has been recently realized that certain oscillator heterogeneities can stabilize a synchronization orbit of the homogeneous system (23–25). In these studies, the heterogeneity is purposively designed to preserve at least one common orbit among the nonidentical oscillators, which may not always be practical to achieve in applications.
Here, we show that oftentimes random differences among individual oscillators can consistently stabilize the dynamics around an otherwise unstable synchronization orbit of the homogeneous system. We demonstrate the phenomenon for random heterogeneity in delay-coupled Stuart–Landau oscillators, which is a canonical model for limit-cycle oscillations close to a Hopf bifurcation. Stuart–Landau oscillators have been used to describe numerous processes, ranging from electrochemical reaction oscillations (26) to plant circadian rhythms (27). Importantly, we establish that, to preserve system-level coherence, random heterogeneity can be more effective than heterogeneity purposely designed to optimize the stability of identical synchronization. To support our theoretical and numerical results, we performed experiments using coupled electrochemical oscillators. The experimental results confirm our predictions and further demonstrate the effect for systems not in the close vicinity of a Hopf bifurcation. These findings are expected to have implications for a broad class of natural and engineered systems, whose functions depend on the synchronization of heterogeneous components.
Results
Modeling the Dynamics of Heterogeneous Oscillators.
We consider a network of delay-coupled nonidentical Stuart–Landau oscillators, whose dynamics are governed bywhere is a complex variable representing the state of the th oscillator and is the coupling strength. The adjacency matrix encodes the network structure, and is the indegree for oscillator . The coupling delay models the finite speed of signal propagation in real systems, which is often significant in biological (28–30), physical (31–33), and control systems (34, 35). The local dynamics are expressed in the following canonical form for systems born out of a Hopf bifurcation (36):where , , and are real parameters associated with the amplitude, base frequency, and amplitude-dependent frequency of the underlying limit-cycle oscillations.
[1]
[2]
The oscillators are identical when , and for all . For identical oscillators, the identical synchronization state is defined by the limit-cycle synchronous solutionThe amplitude and angular velocity can be found by solving the transcendental equationswhere (10). When random heterogeneity is introduced through one or more oscillator parameters, the identical synchronization state [3] will, in general, no longer exist. Nonetheless, we show that heterogeneous systems can still admit states that are synchronized in the sense of exhibiting cohesive phase and amplitude dynamics, as formalized below. Here, we consider synchronization in this sense and ask whether it can be stabilized by random oscillator heterogeneity.
[3]
[4a]
[4b]
We start by considering a homogeneous system consisting of identical Stuart–Landau oscillators coupled through a directed ring network ( if and otherwise) for , , , , and . Under this parameter choice, the limit-cycle synchronous solution [3] is unstable, and the system evolves into a symmetry-broken state, exhibiting incoherent chaotic dynamics. As an example of a heterogeneous system, we consider the same network with the base angular velocity drawn from a Gaussian distribution with mean (as in the homogeneous system) and SD . For additional details on the numerical procedure, see Materials and Methods.
In Fig. 1, we show typical trajectories, order parameters, and angular velocities for the homogeneous and heterogeneous systems. Here, two order parameters are introduced to measure the cohesiveness of the dynamics: the phase order parameter , which is the one typically used in the study of Kuramoto oscillators; and the phase-amplitude order parameter , which measures the coherence in both phases and amplitudes. It follows that both and are constant for frequency-synchronized states. For the trajectories shown, the two systems were initialized close to the limit-cycle synchronous state. The homogeneous system loses synchrony at t = 1,500 and transitions to an incoherent state with both and fluctuating around 0.2. Remarkably, despite having different base frequencies, oscillators in the heterogeneous system converge to a stable cohesive state with large order parameters () and identical angular velocities. That is, the oscillators are not only approximately synchronized in phase and amplitude—they are also exactly synchronized in frequency (i.e., phase-locked). For an animation of this phenomenon in larger systems, see Movie S1. It is worth noting that the same phenomenon is also observed when the coupling function is nonlinear and/or when the coupling delay is link-dependent, which we demonstrate in SI Appendix, SI Text and Figs. S1 and S2.
Movie S1.
Animation showing the evolution of four systems of delay-coupled Stuart-Landau oscillators (snapshots shown above for two different times t). Each system consists of 36 oscillators (represented as dots moving in the complex plane), with different levels of random heterogeneity indicated by colors and quantified by σ. All four systems are initialized close to the identical synchronization state. After about 4000 time units, both systems on the diagonal (no heterogeneity and large heterogeneity) have evolved into an incoherent state, while the off-diagonal systems (intermediate heterogeneity) maintain a high level of synchronization, which persists indefinitely. The network structure is a directed ring and the parameters are the same as in Fig. 1.
Fig. 1.

Synchronization States and Stability Conditions.
To gain theoretical understanding of the effect shown in Fig. 1 and its prevalence, we characterize the synchronization states of interest in the presence of heterogeneity and derive analytical conditions for their stability. These results are established for delay-coupled Stuart–Landau oscillators with arbitrary heterogeneity.
Inspired by the fact that the heterogeneous system in Fig. 1B settles into a frequency-synchronized state, for which the frequencies of all oscillators are equal and the phase differences and amplitudes remain constant, we employ the following ansatz:where oscillator has amplitude and phase lag , both of which are assumed to be constant, and all oscillators share the same angular velocity . Substituting this ansatz into Eq. 1, we obtain nonlinear algebraic equations with unknowns:for , where . Taking , which can be done without loss of generality, the solution of Eqs. 6a and 6b determines , and . As shown in Fig. 2, when parameter heterogeneity is not too large, this gives us frequency-synchronized states that are close to the identical synchronization state of the homogeneous system given by Eqs. 3 and 4.
[5]
[6a]
[6b]
Fig. 2.

We can analyze the stability of the frequency-synchronized states through the variational equation that governs the evolution of small deviations and from those states. Taking , , and , the variational equation can be written aswhere , , , , and . Here, is a matrix obtained by replacing each entry in with a block given by .
[7]
Since all of the matrices in Eq. 7 are time-independent, we can assume and reduce the stability calculation to determining the exponents that solve the following characteristic equation:where indexes the solutions. The Lyapunov exponents of Eq. 7 are given by the real parts of . One Lyapunov exponent (referred to as ) is identically null and corresponds to the perturbation mode parallel to the frequency-synchronization manifold (in which the phases of all oscillators are subject to the same perturbation). The other Lyapunov exponents correspond to perturbation modes transverse to the synchronization manifold. The stability of the frequency-synchronization state is determined by the sign of the maximum transverse Lyapunov exponent (MTLE), which is given by .
[8]
Disorder Consistently Promotes Synchronization.
We now examine systematically the phenomenon of synchronization induced by random heterogeneity. In particular, we address key questions underlying its prevalence. For example, what is the effect of the magnitude of parameter mismatches? Do the results change significantly depending on which parameters are made heterogeneous? Most importantly, can different realizations of random heterogeneity consistently induce synchronization?
In Fig. 3, we start with the same homogeneous system as in Fig. 1A and introduce heterogeneity in , , and , respectively. (Here, the factor is introduced to scale for , because the influence of in Eq. 2 is scaled by the square of the oscillation amplitude. The constant can be found by solving Eqs. 4a and 4b for the corresponding homogeneous system.) In all cases, the SD is , and the mean is taken to be the same as the corresponding parameter in the homogeneous system. For each realization of heterogeneity, as increases from zero, the identical synchronization state progressively changes into a phase-locked state with large order parameters. The stability of this state is measured by , which we obtain by solving Eq. 8 for each realization of heterogeneity. The filled green curves in Fig. 3 A–C, Upper, show the probability that synchronization is stabilized by random heterogeneity in each parameter. These results are verified by direct simulations of Eqs. 1 and 2 for various , shown as purple circles. In Fig. 3 A–C, Lower, we plot for a representative subset of realizations of heterogeneity in each parameter, visualizing their impact on stability as an ensemble.
Fig. 3.

One can see from Fig. 3 that there is always a sweet spot of optimal heterogeneity at an intermediate value of . Around that sweet spot, the green curves stay very close to one, indicating that intermediate levels of heterogeneity can consistently induce synchronization, largely independent of its particular realization. It is interesting to note from Fig. 3 A–C, Lower, that small heterogeneity always improves stability under the conditions considered, as reflected in the monotonic decrease of for small . Disorder can also consistently stabilize synchronization when all three parameters are allowed to be heterogeneous, as demonstrated in SI Appendix, SI Text and Fig. S3. Furthermore, we verified that the same effect can be observed for a wide range of network sizes and different network structures (SI Appendix, SI Text and Figs. S4–S6).
Disorder Can Be Better than Design.
It is important to compare the effect of random and nonrandom heterogeneities. When the heterogeneity is purposively designed, Stuart–Landau oscillators can synchronize identically (i.e., all phase differences are identically zero and all amplitudes are equal), even though they are nonidentical. This is most easily seen from Eqs. 4a and 4b, whose solution remains invariant under the transformation , for any . Thus, any given Stuart–Landau oscillator belongs to a continuous family of nonidentical Stuart–Landau oscillators parameterized by , within which the oscillators can synchronize identically with each other. Moreover, as shown in ref. 25, mixing different oscillators from the same family can stabilize identical synchronization that would otherwise be unstable.
By designing heterogeneity to preserve identical synchronization, can we do better than by relying on random heterogeneity? Once again, we start with the homogeneous system studied in Figs. 1 and 3. The oscillators are then made heterogeneous by sampling from the identically synchronizable family, with drawn from a Gaussian distribution. More concretely, and , where has SD and mean zero. This can be seen as a special subset of oscillators with random heterogeneity in both parameters and , the crucial difference being that and are not independent for the designed heterogeneity.
In Fig. 4, we compare the ensemble average MTLE and order parameters between systems with random heterogeneity and systems with designed heterogeneity on the same parameters. Consistent with Fig. 3, random heterogeneity is most effective for intermediate magnitudes , ranging from 0.05 to 0.1. On the other hand, designed heterogeneity is effective for much larger , from about 0.4 to 0.6, which may be interpreted as a consequence of the identical synchronization solution being preserved in this case. Remarkably, no system with designed heterogeneity is stable within the range for which random heterogeneity is effective. This implies that at intermediate magnitude, random heterogeneity can outperform heterogeneity specifically designed to preserve identical synchronization.
Fig. 4.

Insight from a Minimal System.
To gain further understanding, in Fig. 5, we focus on a minimal system formed by three nonidentical Stuart–Landau oscillators coupled through a directed ring network. The th oscillator has parameters , with the constraint that . The parameter is introduced to vary the synchronization stability of the homogeneous system without altering the synchronous solution. This enables us to investigate all possible realizations of heterogeneous for different levels of instability by sweeping the – plane.
Fig. 5.

In Fig. 5 A and B, the origin is the only point corresponding to a homogeneous system, and the differences among oscillators increase as one moves away from the origin along the radial directions. Stability analysis indicates that regions of stability appear for intermediate levels of oscillator heterogeneity, as shown in Fig. 5A (, blue belts). The phase-locked state is unstable for weak disorder (, red areas) and ceases to exist for strong disorder (blank areas). A complementary perspective is offered by direct simulations, as shown in Fig. 5B. Because order parameters averaged over time are a poor indicator of coherence for systems with a small number of oscillators, we quantify the level of coherence using the minimum of over a period of 10,000 time units after the initial transient. For zero and small heterogeneity, the three oscillators are in an incoherent state, with . As is increased further, the oscillators first settle into an approximate synchronization state with ranging from 0.6 to 0.9 (light purple regions). The level of coherence continues to improve until it plateaus at for phase-locked states (dark purple regions), which correspond to the stable states marked by the blue belts in Fig. 5A. Finally, once we cross the outer boundary, synchrony is lost again, and the value of falls back to approximately zero. This incoherence–coherence–incoherence transition is illustrated in Fig. 5C with representative trajectories from each stage. It is worth noting that even before the phase-locked state is fully stabilized, disorder can already induce approximate synchronization states with well-defined rhythms, as illustrated by the second trajectory.
Fig. 5 demonstrates two competing effects of disorder: When heterogeneity is too small, it cannot tame synchronization instability; when it is too large, it destroys the synchronization state. In other words, there is a trade-off between synchronizability and stability, and stable synchronization naturally emerges at intermediate levels of disorder. Another interesting observation is that the stable belts are contiguous in all cases in Fig. 5A and completely surround the unstable regions in the middle, which explains why intermediate levels of heterogeneity can consistently stabilize synchronization. It also demonstrates that the effect is robust against increasing instability (controlled by ) in the homogeneous system.
Electrochemical Experiments.
A natural question at this point is whether the described phenomenon is robust and general enough to be observed in real systems. To provide an answer, we performed experiments using chemical oscillators based on the electrochemical dissolution of nickel in sulfuric acidic medium (26). The experimental apparatus consists of a counter electrode, a reference electrode, a potentiostat, and nickel wires submerged in the same sulfuric acidic medium, each attached to a resistor (Fig. 6A). At constant circuit potential ( V relative to the reference electrode) and with the resistance of resistors set to kohm, the dissolution rate of each nickel wire, measured as its current, exhibits periodic oscillations (37). The oscillatory dynamics originate from a Hopf bifurcation at V. Compared to the circuit potential used in some previous studies (33), the system here is farther away from the bifurcation point. When the wires are placed sufficiently far from each other, the current oscillations do not show noticeable synchronization, confirming that the interactions through the solution are negligible. Coupling among the wires can be introduced through external feedback (26, 33), in which the circuit potentials of the wires are set based on the measured currents asfor , where and are the experimental coupling strength and delay, respectively. Here, we investigate wires with oscillatory currents arranged in an undirected four-by-four lattice network with periodic boundary conditions, which can be seen as a two-dimensional variant of the ring networks considered above. For additional details on the experimental setup and procedure, see Materials and Methods.
[9]
Fig. 6.

For relatively strong coupling ( V/mA) and no delay ( s), the system exhibits in-phase synchronization (38). Similar in-phase synchronization exists for large delay ( s) that corresponds to the mean period of the uncoupled oscillations. When is set to s (about half of the oscillation period), the system exhibits a two-cluster state in which every other element on the grid is in phase, and the neighboring elements are in antiphase. When the delay is set between these two regions ( s), the system exhibits a desynchronized state. Nominal oscillator heterogeneity was introduced by setting the resistance of each oscillator to a different value , while keeping the mean resistance fixed to kohm. The level of nominal heterogeneity is measured by the SD among all .
First, we randomly picked one realization of heterogeneity and experimentally tested its effect on the collective dynamics at different levels of heterogeneity . Each experiment was initiated close to the in-phase synchronization state and consisted of running the system for 600 s. The level of coherence was measured by the synchronization error e, defined as the SD among the currents at time . In this case, the synchronization error is a more natural measure of coherence than order parameters because the experimental system is not in the close vicinity of a Hopf bifurcation, and the dynamics of the amplitude variables are oscillatory. (Nevertheless, we verified that the order parameters of the phases extracted using either Hilbert transform or peak-detection algorithms give similar results as the ones obtained using the synchronization error.) The experimental results summarized in Fig. 6B reveal a well-defined minimum of the average synchronization error ⟨e⟩ (averaged over the last 200 s of each experimental run) for an intermediate level of nominal heterogeneity, kohm. This optimization of synchronization at intermediate levels of heterogeneity is consistent with what we observed numerically for delay-coupled Stuart–Landau oscillators.
Unlike the idealized systems used in simulations, experimental systems come with unavoidable imperfections and uncertainties. As a result, the electrochemical oscillators in our experiments have slightly different dynamics, even when the resistances are all set to the same nominal value. These relatively small inherent heterogeneities can arise because of unavoidable differences in the metal wires (e.g., in composition and size) and surface conditions (oxide film layer thickness, localized corrosion, etc.). To account for such inherent heterogeneity, we use peak-detection algorithms (39) to extract the natural frequency and amplitude of each uncoupled oscillator, and we use that information to calculate the measured oscillator heterogeneity for both systems with homogeneous and systems with heterogeneous . Here, , where () is the SD of the oscillation periods (amplitudes) of the uncoupled oscillators normalized by the mean. (For additional details on the data-analysis protocol, see Materials and Methods.) In Fig. 6C, we show results for five sets of independent experiments. Each experiment corresponds to a different realization of heterogeneous resistances (for fixed at 0.13 kohm) and of the homogeneous system (corresponding to kohm). It can be seen that, when uncoupled, all heterogeneous systems have a much higher measured oscillator heterogeneity than the homogeneous systems. In contrast, when coupled, the heterogeneous systems achieve significantly better coherence than the homogeneous systems, which is reflected by a consistently smaller ⟨e⟩.
The striking difference between the behavior of the homogeneous and heterogeneous systems is further visualized in Fig. 7. There, we compare the dynamics in the first ( kohm) and the fifth ( kohm) data points from Fig. 6B. The time series of the homogeneous system (Fig. 7A) is very much incoherent compared to that of the heterogeneous system (Fig. 7B). Accordingly, the heterogeneous system exhibits a smaller synchronization error and a more regular rhythm, as shown in Fig. 7 C and D. In SI Appendix, SI Text and Fig. S7, we show that the same phenomenon can be observed when random shortcuts are added to the four-by-four lattice [which creates a small-world network (40)], suggesting that our conclusions extend to networks with heterogeneous degrees.
Fig. 7.

Discussion
It is often challenging, if not impossible, to completely eliminate component mismatches in oscillator networks. Our results suggest that, rather than trying to erase these imperfections (often to no avail), one may instead be able to take advantage of them to promote synchronization required for the system to function. Indeed, our theory, simulations, and experiments consistently show that synchronization can often be stabilized by intermediate levels of random oscillator heterogeneity. The fact that no fine tuning of the heterogeneity profile is needed to induce synchronization can be valuable for stabilizing synchronization in both technological and biological systems. For example, it is often important to generate high-power output of coherent light in laser systems. Semiconductor diode lasers are of interest in many applications due to their low cost, portability, and ease of fabrication, but a single diode laser typically generates an output power of no more than a few watts (41). It is thus desirable to couple many diode lasers together and exploit their frequency synchronization to increase the emission power (42). While in practice, no two lasers are perfectly identical, this study indicates that it might be possible to boost the performance of coupled laser arrays by harnessing, rather than eliminating, the existing mismatches.
In physiology, many important rhythmic processes also depend on the coordination and coherence among a diverse population of cells (43). The heartbeat, for example, is generated by the synchronized activity of thousands of cardiac pacemaker cells in the sinoatrial node (44, 45), whereas the sleep–wake cycle is regulated by the mutual entrainment of circadian cells in the suprachiasmatic nucleus (46–48). Our findings thus raise the question of whether the heterogeneity among pacemaker or circadian cells is a limitation of the biology or, instead, a feature selected for by evolution to promote synchronization and stabilize vital rhythms in living organisms. On the other hand, in situations in which synchronization is undesirable, such as epilepsy (49), the effect demonstrated here can potentially explain why these pathological states appear to be persistent and difficult to suppress, despite the inherent diversity of neuronal populations. This, in turn, might lead to new ideas for therapeutic interventions.
The effect of disorder is also a recurring theme in condensed-matter physics (50). For example, exotic materials such as topological insulators have attracted a vast amount of attention over the past decade (51, 52). A defining property of topological insulators is the existence of edge states that are protected by time-reversal symmetry, which makes the states robust to weak disorder. In the context of oscillator networks, we have been able to go one step further and identify systems and parameter regions for which synchronization is not only immune to disorder, but also enhanced by it.
There are also interesting similarities and differences between the phenomenon described here and noise-induced synchronization (53, 54). It is well established that spatially correlated (e.g., common) noise can facilitate synchronization (55–57), even if the noise is temporally uncorrelated (i.e., white). In contrast, the disorder we consider here is spatially uncorrelated and temporally quenched. Understanding how the spatial and temporal features of disorder and noise influence a system’s collective dynamics has been an ongoing research effort and a source of new insights. For example, it has been shown that quenched disorder can induce coherence resonance in driven bistable systems (20) and that spatially uncorrelated noise can outperform common noise in increasing coherence when oscillators are nonidentical (58). Conversely, it has also been shown that quenched disorder can mitigate desynchronization instabilities caused by noise (59).
Finally, it is instructive to reflect on three salient characteristics of the results established here. First, oscillator heterogeneity can stabilize frequency synchronization states that are similar to the otherwise unstable states observed in the absence of heterogeneity. This is important because it shows that the stability provided by heterogeneity does not come at the price of qualitatively changing the nature of the dynamics. Second, this stabilization is achieved with high success rate using random parameter heterogeneity, making it easy to implement in real-world systems. Third, the effect can be observed for a wide range of network structures, including networks in which all oscillators are identically coupled. The latter is significant because it shows that the stabilizing effect of oscillator heterogeneity is more fundamental than just counterbalancing instabilities that could have been caused by heterogeneities in the network structure. Thus, our results reveal an important avenue through which system disorder can give rise to emergent dynamical order. Future studies further exploring the relation between system disorder and dynamical order will undoubtedly deepen our understanding of collective behavior and of new means to stabilize and control the dynamics of complex systems.
Materials and Methods
Numerical Procedure.
Delay-coupled Stuart–Landau oscillators were simulated by employing the dde23 integrator in MATLAB, with the relative and absolute tolerances both set to . To initialize an oscillator network, we introduced a random perturbation of the order of to the synchronization state at . Each system was then evolved for 10,000 time units, which is long enough for the oscillators to settle into either a coherent state (if synchronization is stable) or an incoherent state (if synchronization is unstable). Our code for simulating delay-coupled Stuart–Landau oscillators can be found at https://github.com/y-z-zhang/disorder_sync.
Experimental Protocols.
The experiments were performed by using a standard three-electrode cell with a platinum counter, a Hg/Hg2SO4/satK2SO4 reference, and a nickel-array working electrode. The electrolyte was 3 M H2SO4 at 10 °C. The electrode array consisted of 16 1-mm-diameter nickel wires with a spacing of 3 mm. The wires were embedded in epoxy, so that only the wire ends were exposed to the electrolyte. Before the experiments, the electrode array was polished with a series of sandpapers. A multichannel potentiostat (catalog no. Gill-IK64, ACM Instruments), interfaced with a real-time LabVIEW controller (60), was used to measure the current and set the potential of the th wire according to Eq. 9 at a rate of 200 Hz. Throughout the experiments, we set the circuit potential to 1.24 V. Without heterogeneity, the individual resistors were set to 1.06 kohm. Heterogeneity was introduced by setting the individual resistors to different nominal values drawn from a normal distribution, while keeping the mean resistance fixed at 1.06 kohm. To avoid accidentally balancing out the inherent heterogeneity, only random realizations of nominal heterogeneity that had a negligible correlation with the natural frequencies of the unperturbed oscillators were used (we required that the absolute value of the correlation coefficient be smaller than 0.2). The coupling delay was set to 75% of the mean natural period of the oscillations, which corresponds to in the range of 1.50 to 1.75 s throughout the experiments. The coupling strength was set to values about 10% larger in magnitude than the desynchronization threshold (between and V/mA in the reported experiments).
Data-Analysis Protocols.
The peak-detection algorithm finds all local maxima by comparing the neighboring values in a time series. The mean of the detected peaks is taken as the oscillation amplitude, and the mean distance between consecutive peaks is the oscillation period. Our data-analysis scripts and experimental data are available at https://github.com/y-z-zhang/disorder_sync. By following the Jupyter Notebooks included in the GitHub repository, one can explore the data interactively and reproduce the results presented in Figs. 6 and 7 and SI Appendix, Fig. S7.
Data Availability
Experimental time-series data have been deposited in GitHub (https://github.com/y-z-zhang/disorder_sync).
Acknowledgments
We thank Eberhard Bodenschatz, Kyoung-Jin Lee, and Yehuda Braiman for insightful discussions. Y.Z. and A.E.M. were supported by Army Research Office Grants W911NF-15-1-0272 and W911NF-19-1-0383. J.L.O.-E. and I.Z.K. were supported by NSF Grant CHE-1900011. J.L.O.-E. also was supported by Consejo Nacional de Ciencia y Tecnología; and Y.Z. was further supported by the Schmidt Science Fellowship.
Supporting Information
Appendix (PDF)
- Download
- 3.15 MB
Movie S1.
Animation showing the evolution of four systems of delay-coupled Stuart-Landau oscillators (snapshots shown above for two different times t). Each system consists of 36 oscillators (represented as dots moving in the complex plane), with different levels of random heterogeneity indicated by colors and quantified by σ. All four systems are initialized close to the identical synchronization state. After about 4000 time units, both systems on the diagonal (no heterogeneity and large heterogeneity) have evolved into an incoherent state, while the off-diagonal systems (intermediate heterogeneity) maintain a high level of synchronization, which persists indefinitely. The network structure is a directed ring and the parameters are the same as in Fig. 1.
- Download
- 11.54 MB
References
1
R. Olfati-Saber, A. Fax, R. M. Murray, Consensus and cooperation in networked multi-agent systems. Proc. IEEE 95, 215–233 (2007).
2
P. D. Manrique, N. F. Johnson, Individual heterogeneity generating explosive system network dynamics. Phys. Rev. E 97, 032311 (2018).
3
S. H. Strogatz, From Kuramoto to Crawford: Exploring the onset of synchronization in populations of coupled oscillators. Physica D 143, 1–20 (2000).
4
A. Arenas, A. Díaz-Guilera, J. Kurths, Y. Moreno, C. Zhou, Synchronization in complex networks. Phys. Rep. 469, 93–153 (2008).
5
D. A. Wiley, S. H. Strogatz, M. Girvan, The size of the sync basin. Chaos 16, 015103 (2006).
6
Z. Li, Z. Duan, G. Chen, L. Huang, Consensus of multiagent systems and synchronization of complex networks: A unified viewpoint. IEEE Trans. Circuits Syst. I, Reg. Papers 57, 213–224 (2010).
7
K. P. O’Keeffe, P. L. Krapivsky, S. H. Strogatz, Synchronization as aggregation: Cluster kinetics of pulse-coupled oscillators. Phys. Rev. Lett. 115, 064101 (2015).
8
L. M. Pecora, T. L. Carroll, Master stability functions for synchronized coupled systems. Phys. Rev. Lett. 80, 2109–2112 (1998).
9
T. Nishikawa, A. E. Motter, Synchronization is optimal in nondiagonalizable networks. Phys. Rev. E 73, 065106 (2006).
10
C. U. Choe, T. Dahms, P. Hövel, E. Schöll, Controlling synchrony by delay coupling in networks: From in-phase to splay and cluster states. Phys. Rev. E 81, 025205 (2010).
11
L. Tang, X. Wu, J. Lü, Ja. Lu, R. M. D’Souza, Master stability functions for complete, intralayer, and interlayer synchronization in multiplex networks of coupled Rössler oscillators. Phys. Rev. E 99, 012304 (2019).
12
J. G. Restrepo, E. Ott, B. R. Hunt, Spatial patterns of desynchronization bursts in networks. Phys. Rev. E 69, 066215 (2004).
13
J. Sun, E. M. Bollt, T. Nishikawa, Master stability functions for coupled nearly identical dynamical systems. Europhys. Lett. 85, 60011 (2009).
14
T. Pereira, J. Eldering, M. Rasmussen, A. Veneziani, Towards a theory for diffusive coupling functions allowing persistent synchronization. Nonlinearity 27, 501–525 (2014).
15
S. Acharyya, R. Amritkar, Synchronization of nearly identical dynamical systems: Size instability. Phys. Rev. E 92, 052902 (2015).
16
F. Sorrentino, L. Pecora, Approximate cluster synchronization in networks with symmetries and parameter mismatches. Chaos 26, 094823 (2016).
17
Y. Braiman, J. F. Lindner, W. L. Ditto, Taming spatiotemporal chaos with disorder. Nature 378, 465–467 (1995).
18
Y. Braiman, W. Ditto, K. Wiesenfeld, M. Spano, Disorder-enhanced synchronization. Phys. Lett. A 206, 54–60 (1995).
19
S. F. Brandt, B. K. Dellen, R. Wessel, Synchronization from disordered driving forces in arrays of coupled oscillators. Phys. Rev. Lett. 96, 034104 (2006).
20
C. J. Tessone, C. R. Mirasso, R. Toral, J. D. Gunton, Diversity-induced resonance. Phys. Rev. Lett. 97, 194101 (2006).
21
C. J. Tessone, A. Scire, R. Toral, P. Colet, Theory of collective firing induced by noise or diversity in excitable media. Phys. Rev. E 75, 016203 (2007).
22
G. Montaseri, M. Meyer-Hermann, Diversity of coupled oscillators can enhance their synchronization. Phys. Rev. E 94, 042213 (2016).
23
T. Nishikawa, A. E. Motter, Symmetric states requiring system asymmetry. Phys. Rev. Lett. 117, 114101 (2016).
24
Y. Zhang, T. Nishikawa, A. E. Motter, Asymmetry-induced synchronization in oscillator networks. Phys. Rev. E 95, 062215 (2017).
25
Y. Zhang, A. E. Motter, Identical synchronization of nonidentical oscillators: When only birds of different feathers flock together. Nonlinearity 30, R1–R23 (2018).
26
W. Zou et al., Restoration of rhythmicity in diffusively coupled dynamical networks. Nat. Commun. 6, 7709 (2015).
27
H. Fukuda, N. Nakamichi, M. Hisatsune, H. Murase, T. Mizuno, Synchronization of plant circadian oscillators with a phase delay effect of the vein network. Phys. Rev. Lett. 99, 098102 (2007).
28
U. Ernst, K. Pawelzik, T. Geisel, Synchronization induced by temporal delays in pulse-coupled oscillators. Phys. Rev. Lett. 74, 1570–1573 (1995).
29
D. Bratsun, D. Volfson, L. S. Tsimring, J. Hasty, Delay-induced stochastic oscillations in gene regulation. Proc. Natl. Acad. Sci. U.S.A. 102, 14593–14598 (2005).
30
R. Vicente, L. L. Gollo, C. R. Mirasso, I. Fischer, G. Pipa, Dynamical relaying can yield zero time lag neuronal synchrony despite long conduction delays. Proc. Natl. Acad. Sci. U.S.A. 105, 17157–17162 (2008).
31
I. Fischer et al., Zero-lag long-range synchronization via dynamical relaying. Phys. Rev. Lett. 97, 123902 (2006).
32
V. Flunkert, O. D’Huys, J. Danckaert, I. Fischer, E. Schöll, Bubbling in delay-coupled lasers. Phys. Rev. E 79, 065201 (2009).
33
C. Bick, M. Sebek, I. Z. Kiss, Robust weak chimeras in oscillator networks with delayed linear and quadratic interactions. Phys. Rev. Lett. 119, 168301 (2017).
34
K. Pyragas, Continuous control of chaos by self-controlling feedback. Phys. Lett. A 170, 421–428 (1992).
35
K. Gu, S. I. Niculescu, Survey on recent results in the stability and control of time-delay systems. J. Dyn. Syst. Meas. Contr. 125, 158–165 (2003).
36
Y. Kuramoto, Chemical Oscillations, Waves, and Turbulence (Springer Series in Synergistics, Springer Science & Business Media, Berlin, Germany, 1984), vol. 19.
37
I. Z. Kiss, W. Wang, J. L. Hudson, Experiments on arrays of globally coupled periodic electrochemical oscillators. J. Phys. Chem. B 103, 11433–11444 (1999).
38
Y. Zhai, I. Z. Kiss, J. L. Hudson, Control of complex dynamics with time-delayed feedback in populations of chemical oscillators: Desynchronization and clustering. Ind. Eng. Chem. Res. 47, 3502–3514 (2008).
39
SciPy, Peak detection algorithms in SciPy (2021). https://docs.scipy.org/doc/scipy/reference/generated/scipy.signal.find_peaks.html. Accessed 5 May 2021.
40
D. J. Watts, S. H. Strogatz, Collective dynamics of ‘small-world’ networks. Nature 393, 440–442 (1998).
41
J. Ohtsubo, Semiconductor Lasers: Stability, Instability and Chaos (Springer Series in Optical Sciences, Springer, Berlin, Germany, 2012), vol. 111.
42
N. Nair, E. Bochove, Y. Braiman, Phase-locking of arrays of weakly coupled semiconductor lasers. Opt. Express 26, 20040–20050 (2018).
43
L. Glass, Synchronization and rhythmic processes in physiology. Nature 410, 277–284 (2001).
44
D. C. Michaels, E. P. Matyas, J. Jalife, Mechanisms of sinoatrial pacemaker synchronization: A new hypothesis. Circ. Res. 61, 704–714 (1987).
45
A. V. Maltsev et al., Synchronization of stochastic Ca2+ release units creates a rhythmic Ca2+ clock in cardiac pacemaker cells. Biophys. J. 100, 271–283 (2011).
46
J. C. Leloup, A. Goldbeter, Toward a detailed computational model for the mammalian circadian clock. Proc. Natl. Acad. Sci. U.S.A. 100, 7051–7056 (2003).
47
D. Gonze, S. Bernard, C. Waltermann, A. Kramer, H. Herzel, Spontaneous synchronization of coupled circadian oscillators. Biophys. J. 89, 120–129 (2005).
48
T. L. To, M. A. Henson, E. D. Herzog, F. J. Doyle, III, A molecular model for intercellular synchronization in the mammalian circadian clock. Biophys. J. 92, 3792–3803 (2007).
49
P. Jiruska et al., Synchronization and desynchronization in epilepsy: Controversies and hypotheses. J. Physiol. 591, 787–797 (2013).
50
R. Nandkishore, D. A. Huse, Many-body localization and thermalization in quantum statistical mechanics. Annu. Rev. Condens. Matter Phys. 6, 15–38 (2015).
51
M. Z. Hasan, C. L. Kane, Colloquium: Topological insulators. Rev. Mod. Phys. 82, 3045 (2010).
52
X. L. Qi, S. C. Zhang, Topological insulators and superconductors. Rev. Mod. Phys. 83, 1057 (2011).
53
C. Zhou, J. Kurths, I. Z. Kiss, J. L. Hudson, Noise-enhanced phase synchronization of chaotic oscillators. Phys. Rev. Lett. 89, 014101 (2002).
54
E. Ullner, J. Buceta, A. Díez-Noguera, J. García-Ojalvo, Noise-induced coherence in multicellular circadian clocks. Biophys. J. 96, 3573–3581 (2009).
55
J. N. Teramae, D. Tanaka, Robustness of the noise-induced phase synchronization in a general class of limit cycle oscillators. Phys. Rev. Lett. 93, 204103 (2004).
56
H. Nakao, K. Arai, Y. Kawamura, Noise-induced synchronization and clustering in ensembles of uncoupled limit-cycle oscillators. Phys. Rev. Lett. 98, 184101 (2007).
57
K. H. Nagai, H. Kori, Noise-induced synchronization of a large population of globally coupled nonidentical oscillators. Phys. Rev. E 81, 065202 (2010).
58
Z. G. Nicolaou, M. Sebek, I. Z. Kiss, A. E. Motter, Coherent dynamics enhanced by uncorrelated noise. Phys. Rev. Lett. 125, 094101 (2020).
59
F. Molnar, T. Nishikawa, A. E. Motter, Network experiment demonstrates converse symmetry breaking. Nat. Phys. 16, 351–356 (2020).
60
R. Bitter, T. Mohiuddin, M. Nawrocki, LabVIEW: Advanced Programming Techniques (CRC Press, Boca Raton, FL, 2017).
Information & Authors
Information
Published in
Classifications
Copyright
© 2021. Published under the PNAS license.
Data Availability
Experimental time-series data have been deposited in GitHub (https://github.com/y-z-zhang/disorder_sync).
Submission history
Published online: May 21, 2021
Published in issue: May 25, 2021
Change history
June 10, 2021: The SI Appendix, Fig. 7, and the article text have been updated. Previous version (May 21, 2021)
Keywords
Acknowledgments
We thank Eberhard Bodenschatz, Kyoung-Jin Lee, and Yehuda Braiman for insightful discussions. Y.Z. and A.E.M. were supported by Army Research Office Grants W911NF-15-1-0272 and W911NF-19-1-0383. J.L.O.-E. and I.Z.K. were supported by NSF Grant CHE-1900011. J.L.O.-E. also was supported by Consejo Nacional de Ciencia y Tecnología; and Y.Z. was further supported by the Schmidt Science Fellowship.
Notes
This article is a PNAS Direct Submission. M.G. is a guest editor invited by the Editorial Board.
Authors
Competing Interests
The authors declare no competing interest.
Metrics & Citations
Metrics
Altmetrics
Citations
Cite this article
Random heterogeneity outperforms design in network synchronization, Proc. Natl. Acad. Sci. U.S.A.
118 (21) e2024299118,
https://doi.org/10.1073/pnas.2024299118
(2021).
Copied!
Copying failed.
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.