Skip to main content
Log in

Spike-Adding Canard Explosion in a Class of Square-Wave Bursters

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

This paper examines a spike-adding bifurcation phenomenon whereby small-amplitude canard cycles transition into large-amplitude bursting oscillations along a single continuous branch in parameter space. We consider a class of three-dimensional singularly perturbed ODEs with two fast variables and one slow variable and singular perturbation parameter \(\varepsilon \ll 1 \) under general assumptions which guarantee such a transition occurs. The primary ingredients include a cubic critical manifold and a saddle homoclinic bifurcation within the associated layer problem. The continuous transition from canard cycles to N-spike bursting oscillations up to \(N\sim {\mathcal {O}}(1/\varepsilon )\) spikes occurs upon varying a single bifurcation parameter on an exponentially thin interval. We construct this transition rigorously using geometric singular perturbation theory; critical to understanding this transition are the existence of canard orbits and slow passage through the saddle homoclinic bifurcation, which are analyzed in detail.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  • Al-Naimee, K., Marino, F., Ciszak, M., Meucci, R., Arecchi, F.T.: Chaotic spiking and incomplete homoclinic scenarios in semiconductor lasers with optoelectronic feedback. New J. Phys. 11(7), 073022 (2009)

    Article  Google Scholar 

  • Carter, P., Sandstede, B.: Fast pulses with oscillatory tails in the FitzHugh–Nagumo system. SIAM J. Math. Anal. 47(5), 3393–3441 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Carter, P., Sandstede, B.: Unpeeling a homoclinic banana in the FitzHugh–Nagumo system. SIAM J. Appl. Dyn. Syst. 17(1), 236–349 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Chay, T.R., Keizer, J.: Minimal model for membrane oscillations in the pancreatic beta-cell. Biophys. J. 42(2), 181–189 (1983)

    Article  Google Scholar 

  • Deng, B.: Homoclinic bifurcations with nonhyperbolic equilibria. SIAM J. Math. Anal. 21(3), 693–720 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  • Desroches, M., Kaper, T.J., Krupa, M.: Mixed-mode bursting oscillations: dynamics created by a slow passage through spike-adding canard explosion in a square-wave burster. Chaos 23(4), 046106 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Desroches, M., Fernández-García, S., Krupa, M.: Canards in a minimal piecewise-linear square-wave burster. Chaos 26(7), 073111 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Desroches, M., Krupa, M., Rodrigues, S.: Spike-adding in parabolic bursters: the role of folded-saddle canards. Physica D 331, 58–70 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Dumortier, F., Roussarie, R.H.: Canard cycles and center manifolds, volume 577. American Mathematical Soc (1996)

  • Eckhaus, W.: Relaxation oscillations including a standard chase on french ducks. In: Asymptotic Analysis II, pp. 449–497. Springer (1983)

  • Fenichel, N.: Persistence and smoothness of invariant manifolds for flows. Indiana Univ. Math. J. 21(3), 193–226 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  • Fenichel, N.: Geometric singular perturbation theory for ordinary differential equations. J. Differ. Equ. 31(1), 53–98 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  • Guckenheimer, J., Kuehn, C.: Computing slow manifolds of saddle type. SIAM J. Appl. Dyn. Syst. 8(3), 854–879 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Hindmarsh, J.L., Rose, R.M.: A model of the nerve impulse using two first-order differential equations. Nature 296(5853), 162 (1982)

    Article  Google Scholar 

  • Hindmarsh, J.L., Rose, R.M.: A model of neuronal bursting using three coupled first order differential equations. Proc. R. Soc. Lond. B 221(1222), 87–102 (1984)

    Article  Google Scholar 

  • Homburg, A.J., Sandstede, B.: Homoclinic and heteroclinic bifurcations in vector fields. Handb. Dyn. Syst. 3, 379–524 (2010)

    Article  MATH  Google Scholar 

  • Kinney, W.M.: An application of Conley index techniques to a model of bursting in excitable membranes. J. Differ. Equ. 162(2), 451–472 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  • Kinney, W.M.: Applying the Conley index to fast-slow systems with one slow variable and an attractor. Rocky Mt. J. Math. 38(4), 1177–1214 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa, M., Sandstede, B., Szmolyan, P.: Fast and slow waves in the FitzHugh–Nagumo equation. J. Differ. Equ. 133(1), 49–97 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa, M., Szmolyan, P.: Extending geometric singular perturbation theory to nonhyperbolic points–fold and canard points in two dimensions. SIAM J. Math. Anal. 33(2), 286–314 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa, M., Szmolyan, P.: Relaxation oscillation and canard explosion. J. Differ. Equ. 174(2), 312–368 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Lee, E., Terman, D.: Uniqueness and stability of periodic bursting solutions. J. Differ. Equ. 158(1), 48–78 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  • Lin, X.-B.: Using Melnikov’s method to solve Silnikov’s problems. Proc. R. Soc. Edinb. Sect. A: Math. 116(3–4), 295–325 (1990)

    Article  MATH  Google Scholar 

  • Linaro, D., Champneys, A., Desroches, M., Storace, M.: Codimension-two homoclinic bifurcations underlying spike adding in the Hindmarsh–Rose burster. SIAM J. Appl. Dyn. Syst. 11(3), 939–962 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Morris, C., Lecar, H.: Voltage oscillations in the barnacle giant muscle fiber. Biophys. J. 35(1), 193–213 (1981)

    Article  Google Scholar 

  • Nowacki, J., Osinga, H.M., Tsaneva-Atanasova, K.: Dynamical systems analysis of spike-adding mechanisms in transient bursts. J. Math. Neurosci. 2(1), 1–28 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Osinga, H.M., Tsaneva-Atanasova, K.T.: Dynamics of plateau bursting depending on the location of its equilibrium. J. Neuroendocrinol. 22(12), 1301–1314 (2010)

    Article  Google Scholar 

  • Plant, R.E., Kim, M.: On the mechanism underlying bursting in the Aplysia abdominal ganglion R15 cell. Math. Biosci. 26(3–4), 357–375 (1975)

    Article  MATH  Google Scholar 

  • Rinzel, J.: Bursting oscillations in an excitable membrane model. In: Ordinary and partial differential equations, pp. 304–316. Springer (1985)

  • Rinzel, J.: A formal classification of bursting mechanisms in excitable systems. In: Mathematical Topics in Population Biology, Morphogenesis and Neurosciences, pp. 267–281. Springer (1987)

  • Rinzel, J., Troy, W.C.: Bursting phenomena in a simplified oregonator flow system model. J. Chem. Phys. 76(4), 1775–1789 (1982)

    Article  MathSciNet  Google Scholar 

  • Rinzel, J., Ermentrout, G.B.: Analysis of neural excitability and oscillations. Methods Neuronal. Model. 135–169 (1989)

  • Ruschel, S., Yanchuk, S.: Chaotic bursting in semiconductor lasers. Chaos Interdiscip. J. Nonlinear Sci. 27(11), 114313 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Schecter, S.: Exchange lemmas 1: Deng’s lemma. J. Differ. Equ. 245(2), 392–410 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Schecter, S.: Exchange lemmas 2: general exchange lemma. J. Differ. Equ. 245(2), 411–441 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Terman, D.: Chaotic spikes arising from a model of bursting in excitable membranes. SIAM J. Appl. Math. 51(5), 1418–1450 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  • Terman, D.: The transition from bursting to continuous spiking in excitable membrane models. J. Nonlinear Sci. 2(2), 135–182 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  • Tsaneva-Atanasova, K., Osinga, H.M., Rieß, T., Sherman, A.: Full system bifurcation analysis of endocrine bursting models. J. Theor. Biol. 264(4), 1133–1146 (2010)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The author gratefully acknowledges support through NSF Grant DMS-2016216 (formerly DMS-1815315).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul Carter.

Additional information

Communicated by Paul Newton.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A Estimates Near the Saddle Homoclinic Point

A Estimates Near the Saddle Homoclinic Point

In this section, we present a proof of Lemma 3.11. We first quote the following result regarding the nature of solutions to the boundary value problem with entry/exit conditions in the sections \( \Sigma ^\mathrm {h}_A, \Sigma ^\mathrm {h}_B\).

Proposition A.1

(Schecter 2008a, Theorem 2.1) Fix \(\Delta >0\) small. There exists \(K_0,\eta >0\) such that the following holds. For any sufficiently small \(\varepsilon >0\), any \(T>0\) and any \(|Y^*|\le \delta _Y\), there exists a solution \((A,B,Y)(\xi ;Y^*,T)\) to (3.32) with \((A,B,Y)(0)\in \Sigma ^\mathrm {h}_A\) and \((A,B,Y)(T)\in \Sigma ^\mathrm {h}_B\) with \(Y(T;Y^*,T)=Y^*\). Furthermore,

$$\begin{aligned} \begin{aligned} |A(\xi ;Y^*,T)|&\le K_0e^{-\eta \xi }\\ |B(\xi ;Y^*,T)|&\le K_0e^{\eta (\xi -T)}\\ |Y(\xi ;Y^*,T)-\Phi (\xi ,Y^*,T)|&\le K_0\varepsilon e^{-\eta T}, \end{aligned} \end{aligned}$$
(A.1)

where \(\Phi (\xi ,Y^*,T)\) denotes the solution of \({\dot{Y}}=\varepsilon G_1(Y,k,\varepsilon )\) satisfying \(Y(T)=Y^*\). The partial derivatives of \((A,B,Y)(\xi ;Y^*,T)\) with respect to \(\xi ,Y^*,T\) up to order r satisfy the same estimates.

Remark A.2

We remark on the appearance of the factor of \(\varepsilon \) appearing in estimates (A.1) for the solution \(Y(\xi ;Y^*,T)\) which is not present in Schecter (2008a, Theorem 2.1). This is due to the fact that the Y-dynamics are of \({\mathcal {O}}(\varepsilon )\), in contrast to the more general case in Schecter (2008a), where there is no small parameter and hence the center dynamics are \({\mathcal {O}}(1)\).

Proof of Lemma 3.11

We use the formulation of Proposition A.1 to prove the estimates on the local map \(\Pi _\mathrm {loc}\). We fix \(\Delta >0\) and assume \(0<\delta _Y,\delta \ll \Delta \) are taken sufficiently small.

For a solution \((A,B,Y)(\xi ; Y^*,T)\) of Proposition A.1, we set \( {\tilde{A}}(Y^*,T):=A(T;Y^*,T)\) and \({\tilde{B}}(Y^*,T):=B(0;Y^*,T)={\mathcal {O}}(e^{-\eta T})\). The map \(\Pi _\mathrm {loc}\) is then determined by

$$\begin{aligned} \begin{aligned} B_\mathrm {loc}(R,Y^*)&= {\tilde{B}}(Y^*,T)\\ Y_\mathrm {loc}(R,Y^*)&= Y(0;Y^*,T). \end{aligned} \end{aligned}$$
(A.2)

where R is defined via the relation \(\Delta R^\rho = \tilde{A}(Y^*,T)\), and the exponent \(\rho \) is as yet to be determined.

Let \(\Phi (\xi ,Y^*,T)\) denote the solution of \({\dot{Y}}=\varepsilon G_1(Y,k,\varepsilon )\) satisfying \(Y(T)=Y^*\); in particular, \(\Phi (\xi ,Y^*,T)\) satisfies the integral equation

$$\begin{aligned} \Phi (\xi ,Y^*,T) =Y^*+\int _T^\xi \varepsilon G_1(\Phi (\xi ,Y^*,T),k,\varepsilon )\mathrm {d}\xi , \end{aligned}$$
(A.3)

and we have the estimates

$$\begin{aligned} \begin{aligned} \Phi (0,Y^*,T)&=Y^*+\varepsilon \gamma T(1+{\mathcal {O}}(\Delta ))\\ \partial _{Y^*} \Phi (\xi ,Y^*,T)&=1+{\mathcal {O}}( \Delta )\\ \partial _T \Phi (\xi ,Y^*,T)&={\mathcal {O}}( \varepsilon ). \end{aligned} \end{aligned}$$
(A.4)

We now define the functions

$$\begin{aligned} \begin{aligned} {\tilde{\alpha }}_0(Y^*,T)&:= \int _0^T F_1(0,0,\Phi (\xi ,Y^*,T),k,\varepsilon )\mathrm {d}\xi \\&=\int _0^T \alpha + {\mathcal {O}}(\Phi ,\varepsilon ) \mathrm {d}\xi \\ {\tilde{\beta }}_0(Y^*,T)&:= \int _0^T F_2(0,0,\Phi (\xi ,Y^*,T),k,\varepsilon )\mathrm {d}\xi \\&=\int _0^T \beta + {\mathcal {O}}(\Phi ,\varepsilon ) \mathrm {d}\xi , \end{aligned} \end{aligned}$$
(A.5)

where

$$\begin{aligned} \begin{aligned} \partial _{Y^*}{\tilde{\alpha }}_0(Y^*,T)&={\mathcal {O}}(T)\\ \partial _T{\tilde{\alpha }}_0(Y^*,T)&= \alpha + {\mathcal {O}}(\Delta )\\ \partial _{Y^*}{\tilde{\beta }}_0(Y^*,T)&={\mathcal {O}}(T)\\ \partial _T{\tilde{\beta }}_0(Y^*,T)&= \beta + {\mathcal {O}}(\Delta ). \end{aligned} \end{aligned}$$
(A.6)

We further define the functions

$$\begin{aligned} \begin{aligned} {\tilde{\alpha }}(Y^*,T)&:= \int _0^T F_1\left( A(\xi ;Y^*,T),B(\xi ;Y^*,T),Y(\xi ;Y^*,T),k,\varepsilon \right) \mathrm {d}\xi \\ {\tilde{\beta }}(Y^*,T)&:= \int _0^T F_2\left( A(\xi ;Y^*,T),B(\xi ;Y^*,T),Y(\xi ;Y^*,T),k,\varepsilon \right) \mathrm {d}\xi . \end{aligned} \end{aligned}$$
(A.7)

We use the estimates in Proposition A.1 combined with directly integrating Eq. (3.32) in reverse time and obtain

$$\begin{aligned} \begin{aligned} {\tilde{A}}&= \Delta \exp \left( -{\tilde{\alpha }}(Y^*,T) \right) \\ {\tilde{B}}&= \Delta \exp \left( -{\tilde{\beta }}(Y^*,T)\right) . \end{aligned} \end{aligned}$$
(A.8)

Using these expressions along with estimates (A.1), we have that

$$\begin{aligned} \begin{aligned} |{\tilde{\alpha }}(Y^*,T)-{\tilde{\alpha }}_0(Y^*,T)|&= {\mathcal {O}}(\Delta )\\ |{\tilde{\beta }}(Y^*,T)-{\tilde{\beta }}_0(Y^*,T)|&= {\mathcal {O}}(\Delta ) \end{aligned} \end{aligned}$$
(A.9)

and the partial derivatives of these expressions with respect to \(Y^*,T\) are also \({\mathcal {O}}(\Delta )\).

The ultimate goal is to express the quantities \({\tilde{B}}\) and \(Y(0;Y^*,T)\) in terms of the quantities \(R, Y^*\), where we define

$$\begin{aligned} \begin{aligned} R&= \left( \frac{{\tilde{A}}}{\Delta }\right) ^{{\tilde{\beta }}_0/{\tilde{\alpha }}_0}. \end{aligned} \end{aligned}$$
(A.10)

To achieve this, we recall (A.8) combined with (A.9)

$$\begin{aligned} \begin{aligned} {\tilde{A}}&=\Delta \exp \left( -\tilde{\alpha }_0(Y^*,T)+ {\mathcal {O}}(\Delta ) \right) \\&=\Delta \exp \left( -\tilde{\alpha }_0(Y^*,T) \right) (1+ {\mathcal {O}}(\Delta )), \end{aligned} \end{aligned}$$
(A.11)

where the derivatives of the \({\mathcal {O}}(\Delta )\) remainder terms with respect to \(Y^*,T\) are also \({\mathcal {O}}(\Delta )\). Hence,

$$\begin{aligned} \begin{aligned} R&= \left( \frac{{\tilde{A}}}{\Delta }\right) ^{{\tilde{\beta }}_0/{\tilde{\alpha }}_0}\\&=\exp \left( -\tilde{\beta }_0(Y^*,T) \right) (1+ {\mathcal {O}}(\Delta ))^{{\tilde{\beta }}_0/{\tilde{\alpha }}_0}. \end{aligned} \end{aligned}$$
(A.12)

This relation can be used to solve for \(T=T(R,Y^*)\), obtaining

$$\begin{aligned} \begin{aligned} T(R,Y^*)&=-\frac{\log R}{\beta }(1+ {\mathcal {O}}(\Delta )). \end{aligned} \end{aligned}$$
(A.13)

Note, due to the exponent \({{\tilde{\beta }}_0/{\tilde{\alpha }}_0}\) appearing in the remainder term of (A.12), the derivatives of the remainder terms in (A.13) with respect to \(R,Y^*\) no longer satisfy the same estimates. However, we are still able to estimate the first order partial derivatives

$$\begin{aligned} \begin{aligned} \partial _RT(R,Y^*)&=-\frac{1}{\beta R}(1+ {\mathcal {O}}(\Delta ))\\ \partial _{Y^*}T(R,Y^*)&={\mathcal {O}}(\log R), \end{aligned} \end{aligned}$$
(A.14)

by implicitly differentiating (A.12).

We set \(B_\mathrm {loc}(R,Y^*):={\tilde{B}}\) and determine

$$\begin{aligned} \begin{aligned} {\tilde{B}}&=\Delta \exp \left( -\tilde{\beta }_0(Y^*,T) \right) (1+ {\mathcal {O}}(\Delta ))\\&= \Delta R(1+ {\mathcal {O}}(\Delta )), \end{aligned} \end{aligned}$$
(A.15)

where the derivatives of the \({\mathcal {O}}(\Delta )\) remainder terms with respect to \(Y^*,T\) are also \({\mathcal {O}}(\Delta )\), and using the expressions (A.14), we obtain

$$\begin{aligned} \begin{aligned} \partial _RB_\mathrm {loc}(R,Y^*)&= \Delta (1+ {\mathcal {O}}(\Delta ))\\ \partial _{Y^*}B_\mathrm {loc}(R,Y^*)&={\mathcal {O}}(R\log R). \end{aligned} \end{aligned}$$
(A.16)

Next, using (A.1), (A.4), and (A.14), and setting \(Y_\mathrm {loc}(R,Y^*):=Y(0)\), we have that

$$\begin{aligned} \begin{aligned} Y_\mathrm {loc}(R,Y^*)&= Y^*-\frac{\varepsilon \gamma \log R}{\beta }\left( 1+{\mathcal {O}}(\Delta )\right) \\ \partial _RY_\mathrm {loc}(R,Y^*)&= -\frac{\varepsilon \gamma }{\beta R}\left( 1+{\mathcal {O}}(\Delta )\right) \\ \partial _{Y^*}Y_\mathrm {loc}(R,Y^*)&= 1+{\mathcal {O}}(\Delta ). \end{aligned}\end{aligned}$$
(A.17)

Finally, we define \(\rho (R,Y^*):={\tilde{\alpha }}_0/{\tilde{\beta }}_0\), and using (A.5) and (A.14), we have

$$\begin{aligned} \begin{aligned} \rho (R,Y^*)&= \alpha /\beta +{\mathcal {O}}(\Delta )\\ \partial _R\rho (R,Y^*)&= {\mathcal {O}}\left( \frac{\Delta }{R\log R}\right) \\ \partial _{Y^*}\rho (R,Y^*)&= {\mathcal {O}}(1), \end{aligned} \end{aligned}$$
(A.18)

which completes the proof of estimates (3.41). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Carter, P. Spike-Adding Canard Explosion in a Class of Square-Wave Bursters. J Nonlinear Sci 30, 2613–2669 (2020). https://doi.org/10.1007/s00332-020-09631-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-020-09631-y

Keywords

Mathematics Subject Classification

Navigation