Skip to main content
Log in

Population Dynamics of Globally Coupled Degrade-and-Fire Oscillators

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

This paper reports the analysis of a model of pulse-coupled oscillators with global inhibitory coupling, inspired by experiments on colonies of bacteria-embedded synthetic genetic circuits. Populations are represented by one-dimensional profiles and their time evolution is governed by a singular differential equation. We address the initial value problem and characterize the dynamics’ main features. In particular, we prove that all trajectory behaviors are asymptotically periodic, with asymptotic features only depending on the population cluster distribution and on the model parameters. A criterion is obtained for the existence of attracting periodic orbits, which reveals the existence of a sharp transition as the coupling parameter is increased. The transition separates a regime where any cluster distribution can be obtained in the limit of large times, to a situation where only trajectories with sufficiently large groups of synchronized oscillators perdure.

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.

Similar content being viewed by others

Notes

  1. We use the notation \(u(x,t-0):=\lim \limits _{s\rightarrow t,s<t}u(x,s)\). A similar definition holds for \(u(x,t+0)\). Moreover, the sign symbol \(\text {Sgn}\) is defined on \(\mathbb {R}^+\) by

    $$\begin{aligned} \text {Sgn}(u)=1\ \text {if}\ u>0\quad \text {and}\quad \text {Sgn}(0)=0. \end{aligned}$$
  2. Namely, if the relation

    $$\begin{aligned} v(x,t)=u(x,t) ,\ \forall x\ne x_{1},x_{2} \quad \text {and}\quad \left\{ \begin{array}{l} v(x_{1},t)=u(x_{2},t)\\ v(x_{2},t)=u(x_{1},t) \end{array}\right. \end{aligned}$$

    holds for \(t=0\), and if \(u(x,t)\) is a solution of the Eq. (1), then \(v(x,t)\) is also a solution and the previous relation holds for all \(t\in (0,+\infty )\).

  3. \(u^+:=\max \{u,0\}\).

  4. Indeed, we have \(T_{2}u(x)>T_{1}u(x)\) for all \(x\in (0,1]\).

  5. If \(u,v\) are two real functions, then \(u\leqslant v\) (resp. \(u<v\)) means \(u(x)\leqslant v(x)\) (resp. \(u(x)<v(x)\)) for all \(x\in (0,1]\).

  6. If the length distribution period \(N_\text {per}\), defined by

    $$\begin{aligned} N_\text {per}=\min \left\{ k\ :\ \ell _{n+k\ \text {mod}\ N}=\ell _{n},\ \forall n\in \{1,\ldots ,N\}\right\} , \end{aligned}$$

    happens to be smaller than \(N\), it actually suffices to consider the composed map \(F_{R^{N_\text {per}-1}\ell }\circ \cdots \circ F_{R\ell }\circ F_{\ell }\), because appropriate permutations of the profiles associated to iterates of this map are non-increasing finite step functions with identical step length distribution as the initial profile. The same consideration suggests to consider, given any step length distribution, the permutation that minimises the period \(N_\text {per}\).

  7. Notice that the lower trace can be alternatively defined as \(u_\text {inf}^{-1}\circ u\) where the generalized inverse (also called the quantile function in Probability Theory) \(u_\text {inf}^{-1}\) can be defined as

    $$\begin{aligned} u_\text {inf}^{-1}=\inf \left\{ y\in (0,1]\ :\ u(y)\geqslant x\right\} ,\ \forall x\in (0,1]. \end{aligned}$$

    In this viewpoint, the property in this item reads \(u_\text {inf}^{-1}\circ u=\text {Id}\) for every strictly increasing function \(u\). A similar comment applies to the upper trace.

References

  1. Abbott, L.F., van Vreeswijk, C.: Asynchronous states in networks of pulse-coupled oscillators. Phys. Rev. E 48, 1483–1490 (1993)

    Article  Google Scholar 

  2. Acebron, J.A., Bonilla, L.L., Perez-Vicente, C.J., Ritort, F., Spigler, R.: The Kuramoto model: a simple paradigm for synchronization phenomena. Rev. Mod. Phys. 77, 137–185 (2005)

    Article  Google Scholar 

  3. Berger, M.A., Wang, Y.: Bounded semigroups of matrices. Linear Algebra Appl. 166, 21–27 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bressloff, P.C.: Mean-field theory of globally coupled integrate-and-fire neural oscillators with dynamic synapses. Phys. Rev. E 60, 2180–2190 (1999)

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

  8. 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)

    Article  MathSciNet  MATH  Google Scholar 

  9. Jury, E.I.: Theory and applications of the z-transform method. Wiley, New York (1964)

    Google Scholar 

  10. Key, E.S., Volkmer, H.: Private Communication

  11. Kuramoto, Y.: Chemical Oscillations, Waves and Turbulence. Springer, Berlin (1984)

    Book  MATH  Google Scholar 

  12. Kuramoto, Y.: Collective synchronization of pulse-coupled oscillators and excitable units. Physica D 50, 15–30 (1991)

    Article  MATH  Google Scholar 

  13. Mather, W., Bennet, M.R., Hasty, J., Tsimring, L.S.: Delay-induced degrade-and-fire oscillations in small genetic circuits. Phys. Rev. Lett. 102, 068105 (2009)

    Article  Google Scholar 

  14. Mauroy, A., Sepulchre, R.: Global analysis of a continuum model for monotone pulse-coupled oscillators. IEEE Trans. Autom. Control 58, 1154–1166 (2013)

    Article  MathSciNet  Google Scholar 

  15. Mirollo, R.E., Strogatz, S.H.: Synchronization of pulse-coupled biological oscillators. SIAM J. Appl. Math. 50, 1645–1662 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  16. Peskin, C.S.: Mathematical Aspects of Heart Physiology. Courant Institute of Mathematical Science Publication, New York (1975)

    MATH  Google Scholar 

  17. Pikovsky, A., Rosenblum, M., Kurths, J.: Synchronization: A Universal Concept in Nonlinear Sciences. Cambridge University Press, Cambridge (2001)

    Book  MATH  Google Scholar 

  18. Rota, G.-C., Strang, G.: A note on the joint spectral radius. Proc. Neth. Acad. 22, 379–381 (1960)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  21. Thomas, R., D’Ari, R.: Biological Feedbacks. CRC Press, Boca Raton (1990)

    MATH  Google Scholar 

  22. Tyson, J.J., Chen, K., Novak, B.: Network dynamics and cell physiology. Nat. Rev. Mol. Cell Biol. 2, 908–916 (2001)

    Article  Google Scholar 

Download references

Acknowledgments

We are grateful to E. Key and H. Volkmer for promptly providing us a proof of Lemma 5.5.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bastien Fernandez.

Additional information

Bastien Fernandez—On leave from Centre de Physique Théorique.

Appendices

Appendix 1: Discontinuous Dependence on Initial Conditions

Instantaneous resetting simplifies the analysis of  Eq. (1). However, it makes the proof of global existence of solutions rather delicate (and apparently unaccessible by standard approaches such as the Picard operator). It also implies that the solution dependence on initial profiles has discontinuities. In particular, this is the case for the first firing time function \(T_{1}u\).

Indeed, there are examples of sequences \(\{u_n\}\) of profiles that uniformly converge to a limit profile \(u_\infty \) and for which we have \(\lim \limits _{n\rightarrow +\infty }T_{1}u_n(x)\ne T_{1}u_\infty (x)\) for some \(x\in (0,1]\).

To see this, let \(\epsilon \leqslant 1\), let \(u_\infty \) be a profile with a left plateau, i.e. \(u_\infty (x)=u_\infty (x_{1})\) for all \(x\in (0,x_{1}]\) (\(x_{1}>0\)), and let an approximating sequence be defined by

$$\begin{aligned} u_n(x)=\left\{ \begin{array}{lll} u_\infty (x)-\frac{1}{n}&{}\text {if}&{}0<x\leqslant \frac{x_{1}}{2}\\ u_\infty (x)&{}\text {if}&{}\frac{x_{1}}{2}<x\leqslant 1 \end{array}\right. \end{aligned}$$

We obviously have \(T_{1}u_\infty (x)=T_{1}u_\infty (x_{1})\) for all \(x\in (0,x_{1}]\) and direct calculations yield the following result:

$$\begin{aligned} \lim _{n\rightarrow +\infty }T_{1}u_n(x)=\left\{ \begin{array}{l} T_{1}u_\infty (x),\ \forall x\in (0,\frac{x_{1}}{2}]\\ T_{1}u_\infty (x)+\epsilon \eta \frac{x_{1}}{2}(T_{1}u_\infty (x_{1})-u_\infty (x_{1})+1),\ \ \forall x\in (\frac{x_{1}}{2},x_{1}] \end{array}\right. \end{aligned}$$

and the inequalities \(0\leqslant u_\infty (x_{1})-T_{1}u_\infty (x_{1})<1\) imply \(\lim \limits _{n\rightarrow +\infty }T_{1}u_n(x)> T_{1}u_\infty (x)\) for all \(x\in (\frac{x_{1}}{2},x_{1}]\).

In addition, discontinuities may also result in the existence of attracting ghost orbits, depending on parameters. Ghost orbits are periodic cycles of profiles, viz. \(\{u(x,t)\}_{(x,t)\in (0,1]\times \mathbb {R}^+}\) with \(u(\cdot ,t+\tau +0)=u(\cdot ,t)\) for some \(\tau >0\), which, while they do not satisfy Eq. (1), attract all trajectories in their neighborhood (uniform topology). As shown after Theorem 5.3, ghost orbits exist at bifurcation points in the parameter space, when a periodic orbit collapses.

Appendix 2: The Lower and Upper Traces of a Non-decreasing Function

Let \(u:(0,1]\rightarrow (0,1]\) be a left continuous and non-decreasing function. Its lower trace \(\underline{u}\) and respectively upper trace \(\overline{u}\) are defined as follows:

$$\begin{aligned} \underline{u}(x)=\inf \left\{ y\in (0,1]\ :\ u(y)\geqslant u(x)\right\} ,\ \forall x\in (0,1], \end{aligned}$$

and

$$\begin{aligned} \overline{u}(x)=\sup \left\{ y\in (0,1]\ :\ u(y)\leqslant u(x)\right\} ,\ \forall x\in (0,1]. \end{aligned}$$

These functions satisfy the following basic properties.

  • \(0\leqslant \underline{u}(x)\leqslant x\leqslant \overline{u}(x)\leqslant 1\) for all \(x\in (0,1]\).

  • either \(\underline{u}(x)< x\) or \(x<\overline{u}(x)\) implies \(u(y)=u(x)\) for all \(y\in (\underline{u}(x),\overline{u}(x)]\).

  • If \(u\) is strictly increasing, then \(\underline{u}(x)=\overline{u}(x)=x\) for all \(x\in (0,1]\).Footnote 7

  • \(\underline{u}\circ \underline{u}(x)=\underline{u}(x)\) iff \(u\) is continuous at \(x\).

  • \(\overline{u}\circ \overline{u}=\overline{u}\).

  • Both functions \(\underline{u}\) and \(\overline{u}\) are left continuous and non-decreasing. (We prove the property for \(\underline{u}\) here; the proof for \(\overline{u}\) is similar and is left to the reader. Monotonicity is obvious and implies \(\underline{u}(x-0)\leqslant \underline{u}(x)\). Left continuity is also evident in the case \(\underline{u}(x)<x\). If, otherwise \(\underline{u}(x)=x\), there must be a sequence \(\{x_n\}_{n\in \mathbb {N}}\) such that \(u(x_n)<u(x_{n+1})\) and \(\lim \limits _{n\rightarrow +\infty } x_n=x\). The former condition implies \(\underline{u}(x_n)>x_{n-1}\). Together with the latter, we obtain \(\underline{u}(x-0)\geqslant x=\underline{u}(x)\) as desired.)

In our context, the traces provide information about the group structure of a population at time \(t\): \(\underline{u}(x,t)=\overline{u}(x,t)=x\) means that cell \(x\) is isolated, while \(\underline{u}(x,t)<\overline{u}(x,t)\) means that all cells \(y\in (\underline{u}(x,t),\overline{u}(x,t)]\) belong to the same group.

The properties of the lower trace above imply that this function can be entirely determined by its plateaus; namely by considering the following decomposition:

$$\begin{aligned} (0,1]=\mathcal{C}_<\cup \mathcal{C}_=, \end{aligned}$$

where

$$\begin{aligned} \mathcal{C}_<=\left\{ x\in (0,1]\ : \underline{u}(x)<x\right\} \quad \text {and}\quad \mathcal{C}_==\left\{ x\in (0,1]\ : \underline{u}(x)=x\right\} , \end{aligned}$$

the second item above imposes the existence of a countable (possibly empty) set \(\mathcal{D}\) such that \(\mathcal{C}_<=\bigcup \limits _{i\in \mathcal{D}}(x_i^{-},x_i^+]\) where \(x_i^-<x_i^+\leqslant x_{i+1}^-\) for all \(i\). (Notice that \(\mathcal{C}_=\) is empty when \(u\) (or \(\underline{u}\)) is a step function.) In other words, every countable (possibly empty) collection of pairwise disjoint semi-open intervals in \((0,1]\) uniquely defines a lower trace function.

The upper trace function depends only on the lower trace, i.e. \(\overline{u}=\overline{u}\circ \underline{u}\) (and vice-versa, we have \(\underline{u}=\underline{u}\circ \overline{u}\)). One can prove this fact using the sets \(\mathcal{C}_<\) and \(\mathcal{C}_=\) and the analogous decomposition for the upper trace. However, for our purpose, it is more convenient to use the following characterization:

$$\begin{aligned} \overline{u}(x)=\inf \left\{ \underline{u}(y)\ :\ x<\underline{u}(y)\right\} ,\ \forall x\in (0,1], \end{aligned}$$
(13)

with the convention that \(\inf \emptyset =1\) in this expression. To prove this relation, notice first that we must have \(\overline{u}(x)\leqslant \inf \left\{ \underline{u}(y)\ :\ x<\underline{u}(y)\right\} \). Indeed, otherwise there existed \(y\) such that \(x<\underline{u}(y)\) and \(\underline{u}(y)<\overline{u}(x)\). Using that the former inequality is equivalent to \(\overline{u}(x)<y\), it results that we must have

$$\begin{aligned} \underline{u}(y)<\overline{u}(x)<y, \end{aligned}$$
(14)

which is clearly incompatible with the definition of the traces. Secondly, still by using contradiction, assume that \(\overline{u}(x)<\inf \left\{ \underline{u}(y)\ :\ x<\underline{u}(y)\right\} \). This implies the existence of \(z\) such that \(\overline{u}(x)<z<\inf \left\{ \underline{u}(y)\ :\ x<\underline{u}(y)\right\} \). However, the first inequality here implies \(u(x)<u(z)\) and then \(x<\underline{u}(z)\) which contradicts the second inequality.

Similar arguments prove the following relation:

$$\begin{aligned} \overline{u}(x)=\underline{u}(\overline{u}(x)+0)=\inf \{\underline{u}(y)\ :\ \overline{u}(x)<y\},\ \forall x\in (0,\underline{u}(1)). \end{aligned}$$
(15)

Indeed, as before, we must have \(\overline{u}(x)\leqslant \inf \{\underline{u}(y)\ :\ \overline{u}(x)<y\}\) because the converse would yield to the double inequality (14) otherwise. Now if there existed \(z\) such that

$$\begin{aligned} \overline{u}(x)<z< \inf \{\underline{u}(y)\ :\ \overline{u}(x)<y\}, \end{aligned}$$

the right inequality would imply \(z<\underline{u}(y)\leqslant y\) for all \(\overline{u}(x)<y\), hence \(z\leqslant \overline{u}(x)\) holds, which contradicts the left inequality.

In the main text, we also refer to the following relation:

$$\begin{aligned} \int _0^1\underline{u}(x)dx+\int _0^1\overline{u}(x)dx=1. \end{aligned}$$
(16)

In order to prove this relation, consider again the decomposition \(\mathcal{C}_<\cup \mathcal{C}_=\) with \(\mathcal{C}_<=\bigcup \limits _{i\in \mathcal{D}}(x_i^{-},x_i^+]\). A moment’s reflection yields

$$\begin{aligned} \int _0^1\underline{u}(x)dx= & {} \sum _{i\in \mathcal{D}}x_i^-(x_i^+-x_i^-)+\int _\mathcal{C_=}xdx\quad \text {and}\quad \int _0^1\overline{u}(x)dx\\= & {} \sum _{i\in \mathcal{D}}x_i^+(x_i^+-x_i^-)+\int _\mathcal{C_=}xdx, \end{aligned}$$

and then

$$\begin{aligned} \int _0^1\underline{u}(x)dx+\int _0^1\overline{u}(x)dx=\sum _{i\in \mathcal{D}}(x_i^+)^2-(x_i^-)^2+2\int _\mathcal{C_=}xdx. \end{aligned}$$

Equation (16) then directly follows from the fact that

$$\begin{aligned} (x_i^+)^2-(x_i^-)^2=2\int _{x_i^{-}}^{x_i^+}xdx. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Blumenthal, A., Fernandez, B. Population Dynamics of Globally Coupled Degrade-and-Fire Oscillators. J Dyn Diff Equat 29, 523–547 (2017). https://doi.org/10.1007/s10884-015-9449-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10884-015-9449-7

Keywords

Navigation