Skip to main content
Log in

Typical trajectories of coupled degrade-and-fire oscillators: from dispersed populations to massive clustering

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

We consider the dynamics of a piecewise affine system of degrade-and-fire oscillators with global repressive interaction, inspired by experiments on synchronization in colonies of bacteria-embedded genetic circuits. Due to global coupling, if any two oscillators happen to be in the same state at some time, they remain in sync at all subsequent times; thus clusters of synchronized oscillators cannot shrink as a result of the dynamics. Assuming that the system is initiated from random initial configurations of fully dispersed populations (no clusters), we estimate asymptotic cluster sizes as a function of the coupling strength. A sharp transition is proved to exist that separates a weak coupling regime of unclustered populations from a strong coupling phase where clusters of extensive size are formed. Each phenomena occurs with full probability in the thermodynamics limit. Moreover, the maximum number of asymptotic clusters is known to diverge linearly in this limit. In contrast, we show that with positive probability, the number of asymptotic clusters remains bounded, provided that the coupling strength is sufficiently large.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  • Acebron JA, Bonilla LL, Perez-Vicente CJ, Ritort F, Spigler R (2005) The Kuramoto model: a simple paradigm for synchronization phenomena. Rev Mod Phys 77:137–185

    Article  Google Scholar 

  • Bottani S (1995) Pulse-coupled relaxation oscillators: from biological synchronization to self-organized criticality. Phys Rev Lett 74:4189

    Article  Google Scholar 

  • Brunel N, Hakim V (1999) Fast global oscillations in networks of integrate-and-fire neurons with low firing rates. Neural Comput 11:1621–1671

    Article  Google Scholar 

  • Buck J (1988) Synchronous rhythmic flashing of fireflies. II. Q Rev Biol 63:265–289

    Article  Google Scholar 

  • Burkitt AN (2006) A review of the integrate-and-fire neuron model: I. homogeneous synaptic input. Biol Cybern 95:1–19

    Article  MATH  MathSciNet  Google Scholar 

  • Coutinho R, Fernandez B (2004) Fronts in extended systems of bistable maps coupled via convolutions. Nonlinearity 17:23–47

    Article  MATH  MathSciNet  Google Scholar 

  • Danino T, Mondragon-Palomino O, Tsimring LS, Hasty J (2010) A synchronized quorum of genetic clocks. Nature 463:326–330

    Article  Google Scholar 

  • Ernst U, Pawelzik K, Geisel T (1995) Synchronization induced by temporal delays in pulse-coupled oscillators. Phys Rev Lett 74:1570

    Article  Google Scholar 

  • Fernandez B, Tsimring LS (2011) Corepressive interaction and clustering of degrade-and-fire oscillators. Phys Rev E 84:051916

    Article  Google Scholar 

  • Gonze D, Bernard S, Waltermann C, Kramer A, Herzel H (2005) Spontaneous synchronization of coupled circadian oscillators. Biophys J 89:120–129

    Article  Google Scholar 

  • Golomb D, Hansel D, Shraiman B, Sompolinsky H (1992) Clustering in globally coupled phase oscillators. Phys Ref A 45:3516–3530

    Article  Google Scholar 

  • Kolmogorov AN, Fomin SV (1999) Elements of the theory of functions and functional analysis. Dover Publications, Mineaola

    Google Scholar 

  • Lee DeVille RE, Peskin CS, Spencer JH (2010) Dynamics of stochastic neuronal networks and the connections to random graph theory. Math Mod Nat Phenom 5:26–66

    Article  MATH  Google Scholar 

  • Mather W, Bennet MR, Hasty J, Tsimring LS (2009) Delay-induced degrade-and-fire oscillations in small genetic circuits. Phys Rev Lett 102:068105

    Article  Google Scholar 

  • Mirollo RE, Strogatz SH (1990) Synchronization of pulse-coupled biological oscillators. SIAM J Appl Math 50:1645–1662

    Article  MATH  MathSciNet  Google Scholar 

  • Peskin CS (1975) Mathematical aspects of heart physiology. Courant Institute of Mathematical Science Publication, New York

    MATH  Google Scholar 

  • Pikovsky A, Rosenblum M, Kurths J (2003) Synchronization: a universal concept in nonlinear sciences. Cambridge University Press, Cambridge

    Google Scholar 

  • Prindle A, Samayoa P, Razinkov I, Danino T, Tsimring LS, Hasty J (2012) A sensing array of radically coupled genetic ‘biopixels’. Nature 481:39–44

    Article  Google Scholar 

  • Seen W, Urbanczik R (2000) Similar nonleaky integrate-and-fire neurons with instantaneous couplings always synchronize. SIAM J Appl Math 61:1143–1155

    MathSciNet  Google Scholar 

  • Strogatz SH (2000) From Kuramoto to Crawford: exploring the onset of synchronization in populations of coupled oscillators. Physica D 143:1–20

    Article  MATH  MathSciNet  Google Scholar 

  • Tass PA (1999) Phase resetting in medicine and biology. Springer, Berlin

    Book  MATH  Google Scholar 

  • Tsimring LS, Rulkov NF, Larsen ML, Gabbay M (2005) Repulsive synchronization in an array of phase oscillators. Phys Rev Lett 95:14101

    Article  Google Scholar 

  • Van Vreeswijk C, Abbott LF, Ermentrout GB (1994) When inhibition not excitation synchronizes neural firing. J Comput Neurosci 1:313–321

    Google Scholar 

  • Wiesenfeld K, Swift JW (1995) Averaged equations for Josephson junction series arrays. Phys Rev E 51:1020–1025

    Article  Google Scholar 

  • Yamaguchi S, Isejima H, Matsuo T, Okura, K. Yagita R, Kobayashi M, Okamura H (2003) Synchronization of cellular clocks in the suprachiasmatic nucleus. Science’s STKE 302(5649):1408

Download references

Acknowledgments

BF acknowledges stimulating discussions with Jean-Marc Gambaudo and Lai-Sang Young. He is also grateful to Neil Dobbs for pointing out a gap in the original arguments and to the BioCircuits Institute for hospitality during his stay at UCSD. The work of BF was supported by EU Marie Curie fellowship PIOF-GA-2009-235741 and by CNRS PEPS Physique Théorique et ses interfaces and the work of LT was supported by the National Institutes of Health and General Medicine (Grant R01-GM69811) and the San Diego Center for Systems Biology (Grant P50-GM085764).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bastien Fernandez.

Appendices

Appendix A: Mean estimates for configurations in \(\mathcal{T }_N\)

Throughout the proofs, we use the following estimate on the mean \(\frac{1}{N}\sum \nolimits _{k=1}^{N}x_k\) for a subset of configurations \(\{x_k\}_{k=1}^{N-1}\in \mathcal{T }_N\) that has arbitrarily large probability measure. The estimate is a straightforward consequence of the Central Limit Theorem. It can be stated as follows. Recall that the symbol \(\mathbb{P }\) denotes the law of a random variable.

Lemma 7.1

For every \(\delta \in (0,1)\), we have \(\lim \limits _{N\rightarrow \infty }\mathbb{P }(|\frac{1}{N} \sum \nolimits _{k=1}^{N}x_k-\frac{1+\eta }{2}|<\delta )=1\).

Proof

Let \(N\in \mathbb{N },N>1\) be fixed and for every configuration \(x=\{x_k\}_{k=1}^{N-1}\), let \(S_{N-1}(x)=\frac{1}{N-1}\sum \limits _{k=1}^{N-1}x_k\). The quantity \(S_{N-1}\) is regarded as a random variable with law \(\mathbb{P }\).

Consider now the random process in the hypercube \([\eta ,1]^{N-1}\) endowed with the uniform measure \((1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1}\). For this process, the law of \(S_{N-1}\) is simply \((1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}\). A standard argument (presented at the end of this proof below) shows that we have \(\mathbb{P }=(1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}\).

For the process in the hypercube, the quantity \(S_{N-1}\) appears to be the normalized sum of i.i.d. random variables \(x_i\) with Lebesgue distribution in \([\eta ,1]\). The corresponding mean value is \(\frac{1+\eta }{2}\) and the variance is finite. By the Central Limit Theorem, we conclude that for every \(p\in (0,1)\) there exists \(c_p>0\) and \(N_p\in \mathbb{N }\) such that

$$\begin{aligned}&\mathbb{P }\left( \left| S_{N-1}-\frac{1+\eta }{2}\right| \leqslant c_p/\sqrt{N-1}\right) \nonumber \\&\quad \quad =(1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1} \left( \left| S_{N-1}-\frac{1+\eta }{2}\right| \leqslant c_p/\sqrt{N-1}\right) >p,\quad \forall N>N_p. \end{aligned}$$

In particular, for every \(\delta \in (0,1)\), we can ensure that \(|S_{N-1}-\frac{1+\eta }{2}|<\delta /2\) holds with probability larger than \(p\), provided that \(N>\max \{N_p,(2c_p/\delta )^2+1\}\) (so that we also have \(c_p/\sqrt{N-1}< \delta /2\)). Furthermore, the normalization \(x_N=1\) yields the following inequality

$$\begin{aligned} \left| \frac{1}{N}\sum \limits _{k=1}^{N}x_k-\frac{1+\eta }{2}\right| \leqslant \left| S_{N-1}(x)-\frac{1+\eta }{2}\right| +\frac{1}{N}(1-S_{N-1}(x)),\quad \forall x\in \mathcal{T }_N. \end{aligned}$$

By taking \(N>\max \{N_p,(2c_p/\delta )^2+1,2/\delta \}\) (so that we also have \(1/N<\delta /2\)), we can be sure that \(|\frac{1}{N}\sum \nolimits _{k=1}^{N}x_k-\frac{1+\eta }{2}|<\delta \) whenever \(|S_{N-1}-\frac{1+\eta }{2}|<\delta /2\). The Lemma then immediately follows.

It remains to show the equality of laws \(\mathbb{P }=(1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}\). First, notice that we have

$$\begin{aligned} \mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}&=\mathrm{{Leb}}_{N-1}\circ (S_{N-1}|_{C_{N-1}})^{-1}\quad \text{ where }\ C_{N-1}\nonumber \\&=\left\{ x\in [0,1]^{N-1}: i\ne j\Rightarrow x_i\ne x_j\right\} . \end{aligned}$$

Indeed, any subset of \([\eta ,1]^{N-1}\setminus C_{N-1}\) has vanishing \(\mathrm{{Leb}}_{N-1}\) measure. Moreover, we have \(S_{N-1}\circ \sigma =S_{N-1}\) for every permutation of coordinates \(\sigma \). Consequently, the following decomposition holds for every \(\omega \in [\eta ,1]\)

$$\begin{aligned} (S_{N-1}|_{C_{N-1}})^{-1}(\omega )=\bigcup _{\sigma \in \Pi _{N-1}}\sigma \circ (S_{N-1}|_{\mathcal{T }_{N}})^{-1}(\omega ) \end{aligned}$$

where \(\Pi _{N-1}\) is the set of all permutations. By construction, the sets \(\sigma \circ (S_{N-1}|_{\mathcal{T }_{N}})^{-1}(\omega )\) are pairwise disjoints. In addition, they all have the same \(\mathrm{{Leb}}_{N-1}\) measure because permuting coordinates does not affect the volume. Since there are \((N-1)!\) permutations, it results that for every \(\omega \in [\eta ,1]\), we have

$$\begin{aligned} \mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}(\omega )\!=\!(N\!-\!1)!\mathrm{{Leb}}_{N-1}\circ (S_{N-1}|_{\mathcal{T }_{N}})^{-1}(\omega )\!=\!\frac{(N-1)!}{\alpha _N}\mathbb{P }(S_{N-1}=\omega ), \end{aligned}$$

where the last equality follows from the definition of the uniform distribution in Sect. 2. By integrating over \([\eta ,1]\), normalization then implies \(\frac{(N-1)!}{\alpha _N(1-\eta )^{N-1}}=1\), viz. \((1-\eta )^{-(N-1)}\mathrm{{Leb}}_{N-1}\circ S_{N-1}^{-1}=\mathbb{P }\) as desired.\(\square \)

Appendix B: Compactness of the set of increasing functions

Throughout the proofs, we also often need to approximate the piecewise affine interpolation \(\mathrm{{x}}_\mathrm{lin}\) of a configuration \(x\in \mathcal{T }_N\) by a continuous and strictly increasing function chosen in a finite collection. Such approximation relies on the following statement. Let \(\Vert \cdot \Vert _\infty \) denote the uniform norm of a function defined on \([0,1]\).

Proposition 8.1

For every \(\delta >0\), there exists a finite collection \(\{\mathrm{{x}}_{(i,\delta )}\}_{i=1}^{i_\delta }\) of continuous and strictly increasing functions such that, for every piecewise affine continuous increasing function \(\mathrm{{x}}\), there exists \(i\in \{1,\ldots ,i_\delta \}\) such that \(\Vert \mathrm{{x}}-\mathrm{{x}}_{(i,\delta )}\Vert _\infty <\delta \).

This statement is a consequence of a similar property in the weaker \(L^1\)-norm, which we denote by \(\Vert \cdot \Vert _1\).

Lemma 8.2

For every \(\delta >0\), there exists a finite collection \(\{\mathrm{{x}}_{(i,\delta )}\}_{i=1}^{i_\delta }\) of continuous strictly increasing functions such that, for every piecewise affine continuous increasing function \(\mathrm{{x}}\), there exists \(i\in \{1,\ldots ,i_\delta \}\) such that \(\Vert \mathrm{{x}}-\mathrm{{x}}_{(i,\delta )}\Vert _1<\delta \).

Proof of Lemma

By Helly Selection Theorem (Kolmogorov and Fomin 1999), the set of (right continuous) increasing functions from \([0,1]\) into itself is compact for the \(L^1\)-topology. Hence, for every \(\delta >0\), there exists a finite collection \(\{\tilde{\mathrm{{x}}}_{(i,\delta )}\}_{i=1}^{i_\delta }\) of (right continuous) increasing functions such that, for every piecewise affine continuous increasing function \(\mathrm{{x}}\), there exists \(i\in \{1,\ldots ,i_\delta \}\) such that \(\Vert \mathrm{{x}}-\tilde{\mathrm{{x}}}_{(i,\delta )}\Vert _1<\delta /2\).

Let \(h\) be a strictly increasing continuous function from \([-1,1]\) onto \([0,1]\). Then for each extended function \(\tilde{\mathrm{{x}}}_{(i,\delta )}\) on \([-1,1]\) (where \(\tilde{\mathrm{{x}}}_{(i,\delta )}(\omega )=0\) for \(\omega <0\)), consider the function \(\mathrm{{x}}_{(i,\delta )}\) defined by the normalized convolution

$$\begin{aligned} \mathrm{{x}}_{(i,\delta )}(\omega )=\frac{(\tilde{\mathrm{{x}}}_{(i,\delta )}*h)(\omega )}{(\tilde{\mathrm{{x}}}_{(i,\delta )}*h)(1)},\quad \forall \omega \in [0,1] \end{aligned}$$

where \((u*h)(\omega )=\int _{\omega -1}^\omega u(\omega -\theta )~dh(\theta )\) (Lebesgue–Stieltjes integral). Each function \(\mathrm{{x}}_{(i,\delta )}\) is continuous and strictly increasing from \([0,1]\) onto itself. Moreover, by taking \(h\) sufficiently close to the Heaviside function \(H\), one can ensure that \(\Vert \mathrm{{x}}_{(i,\delta )}-\tilde{\mathrm{{x}}}_{(i,\delta )}\Vert _1<\delta /2\) for every \(i\in \{1,\ldots ,i_\delta \}\) and the Lemma follows.

Indeed, if the sequence \(\{h_n\}_{n\in \mathbb{N }}\) pointwise converges to \(H\) on \([-1,1]\), Helly Convergence Theorem (Kolmogorov and Fomin 1999) implies that the sequence \(\{(u*h_n)(\omega )\}_{n\in \mathbb{N }}\) converges to \((u*H)(\omega )=u(\omega )\) for every \(\omega \in [0,1]\). Lebesgue dominated convergence then yields

$$\begin{aligned} \lim \limits _{n\rightarrow \infty }\int _0^1\frac{(u*h_n)(\omega )}{(u*h_n)(1)}~d\omega =\int _0^1 u \end{aligned}$$

from which the desired \(L^1\)-bound on the difference \(\mathrm{{x}}_{(i,\delta )}-\tilde{\mathrm{{x}}}_{(i,\delta )}\) easily follows. \(\square \)

Proof of Proposition 8.1

According to the Lemma, it suffices to show that if \(\{\mathrm{{x}}_n\}_{n\in \mathbb{N }}\) is a sequence of (strictly) increasing functions such that \(\lim \limits _{n\rightarrow \infty }\Vert \mathrm{{x}}-\mathrm{{x}}_n\Vert _1=0\) where \(\mathrm{{x}}\) is continuous, then \(\lim \limits _{n\rightarrow \infty }\Vert \mathrm{{x}}-\mathrm{{x}}_n\Vert _\infty =0\). The proof is similar to that of Lemma B.3 in Coutinho and Fernandez (2004).

By contradiction, assume there exist \(\delta >0\) and a subsequence \(\{\mathrm{{x}}_{n_i}\}_{i\in \mathbb{N }}\) (with \(\lim \limits _{i\rightarrow \infty }n_i=\infty \)) such that

$$\begin{aligned} \Vert \mathrm{{x}}-\mathrm{{x}}_{n_i}\Vert _\infty \geqslant \delta ,\quad \forall i\in \mathbb{N }. \end{aligned}$$

Accordingly, there exists \(\omega _i\in [0,1]\) for every \(i\) such that

$$\begin{aligned} {\text{ either }}\ \mathrm{{x}}(\omega _i)\geqslant \mathrm{{x}}_{n_i}(\omega _i)+\delta \quad \text{ or }\ \mathrm{{x}}(\omega _i)\leqslant \mathrm{{x}}_{n_i}(\omega _i)-\delta . \end{aligned}$$

By taking a subsequence if necessary, we can assume to have either \(\mathrm{{x}}(\omega _i)\geqslant \mathrm{{x}}_{n_i}(\omega _i)+\delta \) for all \(i\in \mathbb{N }\) or \(\mathrm{{x}}(\omega _i)\leqslant \mathrm{{x}}_{n_i}(\omega _i)-\delta \) for all \(i\in \mathbb{N }\).

Assume to be in the first case. The second case can be treated similarly. Since \(\omega _i\in [0,1]\) for all \(i\), there exists a convergent subsequence. W.l.o.g. assume that we have \(\lim \limits _{i\rightarrow \infty }\omega _i=\omega _\infty \).

By compactness, the function \(\mathrm{{x}}\) is uniformly continuous. Let then \(\gamma >0\) be small enough so that we have

$$\begin{aligned} |\mathrm{{x}}(\omega )-\mathrm{{x}}(\omega +\gamma )|<\delta /2,\quad \forall \omega \in [0,1-\gamma ]. \end{aligned}$$

Let now \(\tilde{\omega }\in (\omega _\infty -\delta /2,\omega _\infty )\) be such that \(\lim \limits _{i\rightarrow \infty }\mathrm{{x}}_{n_i}(\tilde{\omega })=\mathrm{{x}}(\tilde{\omega })\). (The existence of \(\tilde{\omega }\) is a consequence of \(L^1\)-convergence.) Convergence to \(\omega _\infty \) and the choice of \(\tilde{\omega }\) imply that we simultaneously have

$$\begin{aligned} |\omega _{i}-\omega _\infty |<\gamma /2\quad \text{ and }\quad \tilde{\omega }<\omega _i,\quad \text{ and } \text{ hence } |\tilde{\omega }-\omega _i|<\gamma , \end{aligned}$$

provided that \(i\) is sufficiently large. The last inequality implies that \(\mathrm{{x}}(\tilde{\omega })-\delta /2\geqslant \mathrm{{x}}(\omega _i)-\delta \) and thus \(\mathrm{{x}}(\tilde{\omega })-\delta /2\geqslant \mathrm{{x}}_{n_i}(\omega _i)\) by the initial assumption. Monotonicity of the \(\mathrm{{x}}_{n_i}\) and the middle inequality above then yield \(\mathrm{{x}}(\tilde{\omega })-\delta /2\geqslant \mathrm{{x}}_{n_i}(\tilde{\omega })\). By taking the limit \(i\rightarrow \infty \), we obtain from the convergence at \(\tilde{\omega }\) that \(-\delta /2\geqslant 0\), which is impossible.

Appendix C: Intensive number of clusters for trajectories starting on equidistant configurations

In this section, we examine the fate at strong coupling, of trajectories initiated from equidistant configurations (or initial conditions close to equidistant configurations) and prove that their asymptotic number of clusters must be intensive. This property is an immediate consequence of the following technical statement.

Lemma 9.1

Let \(\epsilon >\frac{2}{1-\eta }\) and consider the trajectory started from \(x_i=\eta +(1-\eta )\frac{i-1}{N-1}\ (i=1,\ldots ,N)\).

  1. (i)

    For every \(\ell \in \mathbb{N }\) and there exist \(\rho _{\ell }\in (0,1)\) and \(M_{\ell }\in \mathbb{N }\) such that for every \(N>M_{\ell }\), the cluster size \(K_\ell \) at \(\ell \)th firing satisfies \(K_\ell \geqslant \lceil \rho _{\ell } N\rceil \), unless the accumulated reset size \(K_1+\cdots +K_\ell =N\).

  2. (ii)

    We have \(\rho _{\ell +1}>\rho _{\ell }\) for every \(\ell \).

Naturally, property (ii) implies the existence of \(L_\epsilon \) such that \(\sum \limits _{\ell =1}^{L_\epsilon }\rho _\ell \geqslant 1\). Property (i) then forces \(K_1+\cdots +K_{L_\epsilon }=N\) for every \(N>M_{L_\epsilon }\). Thus, for every \(N\in \mathbb{N }\), when starting from the equidistant configuration, the asymptotic number of clusters cannot exceed \(\max \{L_\epsilon ,M_\epsilon \}\).

With a bit of additional effort, one can show that a similar upper bound applies to every trajectory started from configurations in some \(\ell ^\infty \)-neighborhood of the equidistant configuration. (However, this neighborhood has vanishing measure \(\mu \) in the thermodynamics limit.) Therefore, our result indicates that for every \(\epsilon >\frac{2}{1-\eta }\) (a threshold that is larger but close to \(\frac{2}{1+\eta }\)), for every population size, there is positive probability \(\mu \) to obtain an intensive number of clusters in the long time limit.

Proof

We begin by showing the extensive bound on the size \(K_1\) of the first firing cluster. Explicit calculations show that the quantity involved in the definition of \(K_1\) in Sect. 3.2 is given by

$$\begin{aligned} \frac{1}{N}\sum _{k=j+1}^N(x_k-x_j)= \frac{1-\eta }{2}\left( 1-\frac{j}{N}\right) \left( 1- \frac{j-2}{N-1}\right) . \end{aligned}$$

Using \(\frac{j-2}{N-1}<\frac{j}{N}\) yields \(K_1\geqslant \max \{j\in \{1,\ldots ,N\} : (1-\frac{j}{N})^2\geqslant (1-\rho _1)^2\}\) where \(\rho _1\in (0,1)\) is such that \((1-\rho _1)^2=\frac{2}{(1-\eta )\epsilon }\). This quantity \(\rho _1\) exists for every \(\epsilon >\frac{2}{1-\eta }\). It follows that \(K_1\geqslant \lceil \rho _{1} N\rceil \) for all \(N\in \mathbb{N }\) as desired.

For \(\ell >1\), we proceed by induction. Assume that we have already proved that for \(i=1,\ldots ,\ell \), we have \(K_i\geqslant \lceil \rho _{i} N\rceil \) with \(\rho _i\in (0,1)\) provided that \(N\) is sufficiently large. Then, the reasoning at the beginning of the proof of Lemma 5.2 applies here; hence Eq. (9) is a lower bound for \(K_{L+1}\). Using the expression of the equidistant configuration, it easily follows that \(K_{\ell +1}\geqslant \lfloor \frac{\rho _\ell }{1-\eta }(N-1)\rfloor \) (provided that \(K_1+\cdots K_\ell + \lfloor \frac{\rho _\ell }{1-\eta }(N-1)\rfloor \leqslant N\)).

Let then \(M_{\ell +1}\) be sufficiently large so that \(\lfloor \frac{\rho _\ell }{1-\eta }(N-1)\rfloor \geqslant \lceil \frac{\rho _\ell }{1-1.1\eta }N\rceil \) for all \(N>M_{\ell +1}\). Then, we clearly have \(K_{L+1}\geqslant \lceil \rho _{\ell +1} N\rceil \) for all \(N>M_{\ell +1}\) , where \(\rho _{\ell +1}=\frac{\rho _\ell }{1-1.1\eta }>\rho _{\ell }\). The induction follows. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Fernandez, B., Tsimring, L.S. Typical trajectories of coupled degrade-and-fire oscillators: from dispersed populations to massive clustering. J. Math. Biol. 68, 1627–1652 (2014). https://doi.org/10.1007/s00285-013-0680-8

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0680-8

Keywords

Mathematics Subject Classification (2010)

Navigation