Skip to main content
Log in

Dynamics of a Piecewise-Linear Morris–Lecar Model: Bifurcations and Spike Adding

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

Abstract

Multiple-timescale systems often display intricate dynamics, yet of great mathematical interest and well suited to model real-world phenomena such as bursting oscillations. In the present work, we construct a piecewise-linear version of the Morris–Lecar neuron model, denoted PWL-ML, and we thoroughly analyse its bifurcation structure with respect to three main parameters. Then, focusing on the homoclinic connection present in our PWL-ML, we study the slow passage through this connection when augmenting the original system with a slow dynamics for one of the parameters, thereby establishing a simplified framework for this slow-passage phenomenon. Our results show that our model exhibits equivalent behaviours to its smooth counterpart. In particular, we identify canard solutions that are part of spike-adding transitions. Focusing on the one-spike and on the two-spike scenarios, we prove their existence in a more straightforward manner than in the smooth context. In doing so, we present several techniques that are specific to the piecewise-linear framework and with the potential to offer new tools for proving the existence of dynamical objects in a wider context.

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

Similar content being viewed by others

References

  • Arima, N., Okazaki, H., Nakano, H.: A generation mechanism of canards in a piecewise linear system. Trans. IEICE, A 80(3), 447–453 (1997)

    Google Scholar 

  • Baer, S., Erneux, T., Rinzel, J.: The slow passage through a Hopf bifurcation: delay, memory effects, and resonance. SIAM J. Appl. Math. 49(1), 55–71 (1989)

    Article  MathSciNet  Google Scholar 

  • Benoît, E., Callot, J.-L., Diener, F., Diener, M.: Chasse au canard. Collectanea Mathematica 32(1–2), 37–119 (1981)

    MathSciNet  Google Scholar 

  • Bertram, R., Butte, M.J., Kiemel, T., Sherman, A.: Topological and phenomenological classification of bursting oscillations. Bull. Math. Biol. 57(3), 413–439 (1995)

    Article  Google Scholar 

  • Carmona, V., Fernández-García, S., Teruel, A.E.: Saddle-node canard cycles in planar piecewise linear differential systems (2020). arXiv:2003.14112 [math.DS]

  • Carmona, V., Fernández-García, S., Fernández-Sánchez, F., Garcia-Medina, E., Teruel, A.E.: Reversible periodic orbits in a class of 3D continuous piecewise linear systems of differential equations. Nonlinear Anal. Theory Methods Appl. 75(15), 5866–5883 (2012)

    Article  MathSciNet  Google Scholar 

  • Carmona, V., Fernández-García, S., Teruel, A.E.: Saddle-node of limit cycles in planar piecewise linear systems and applications. Discrete Contin. Dyn. Syst. 39(9), 5275–5299 (2019)

    Article  MathSciNet  Google Scholar 

  • Carter, P.: Spike-adding canard explosion in a class of square-wave bursters. J. Nonlinear Sci. (2020). https://doi.org/10.1007/s00332-020-09631-y

    Article  MathSciNet  Google Scholar 

  • Channell, P., Cymbalyuk, G., Shilnikov, A.: Origin of bursting through homoclinic spike adding in a neuron model. Phys. Rev. Lett. 98(13), 134101 (2007)

    Article  Google Scholar 

  • Coombes, S.: Neuronal networks with gap junctions: a study of piecewise linear planar neuron models. SIAM J. Appl. Dyn. Syst. 7(3), 1101–1129 (2008). https://doi.org/10.1137/070707579

    Article  MathSciNet  Google Scholar 

  • Desroches, M., Fernández-García, S., Krupa, M., Prohens, R., Teruel, A.E.: Piecewise-linear (PWL) canard dynamics—simplifying singular perturbation theory in the canard regime using piecewise-linear systems. In: Carmona, V., Cuevas-Maraver, J., Fernández-Sánchez, F., García- Medina, E. (eds.) Nonlinear Systems: Mathematical Theory and Computational Methods, vol. 1, pp. 67–86. Springer, Cham (2018)

    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 Interdiscip. J. Nonlinear Sci. 23(4), 046106 (2013)

    Article  MathSciNet  Google Scholar 

  • Desroches, M., Guillamon, A., Ponce, E., Prohens, R., Rodrigues, S., Teruel, A.E.: Canards, folded nodes, and mixed-mode oscillations in piecewise-linear slow-fast systems. SIAM Rev. 58(4), 653–691 (2016). https://doi.org/10.1137/15M1014528

    Article  MathSciNet  Google Scholar 

  • Desroches, M., Fernández-García, S., Krupa, M.: Canards in a minimal piecewise-linear square-wave burster. Chaos Interdiscip. J. Nonlinear Sci. 26(7), 073111 (2016). https://doi.org/10.1063/1.4958297

    Article  MathSciNet  Google Scholar 

  • Desroches, M., Rinzel, J., Rodrigues, S.: Classification of bursting patterns: a tale of two ducks. PLoS Comput. Biol. 18(2), 1009752 (2022)

    Article  Google Scholar 

  • Di Bernardo, M., Champneys, A., Budd, C., Kowalczyk, P.: Piecewise-Smooth Dynamical Systems: Theory and Applications. Springer, Berlin (2008). https://doi.org/10.1007/978-1-84628-708-4

    Book  Google Scholar 

  • Diminnie, D.C., Haberman, R.: Slow passage through homoclinic orbits for the unfolding of a saddle-center bifurcation and the change in the adiabatic invariant. Physica D 162(1–2), 34–52 (2002)

    Article  MathSciNet  Google Scholar 

  • Dumortier, F., Roussarie, R.: Canard Cycles and Center Manifolds. Memoirs of the American Mathematical Society, vol. 577. American Mathematical Society, Providence (1996)

    Google Scholar 

  • Euzébio, R., Pazim, E.P.R.: Jump bifurcations in some degenerate planar piecewise linear differential systems with three zones. Physica D: Nonlinear Phenomena (2016). https://doi.org/10.1016/j.physd.2016.03.004

    Article  MathSciNet  Google Scholar 

  • Euzébio, R.D., Llibre, J.: On the number of limit cycles in discontinuous piecewise linear differential systems with two pieces separated by a straight line. J. Math. Anal. Appl. 424(1), 475–486 (2015). https://doi.org/10.1016/j.jmaa.2014.10.077

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Fernández-García, S., Desroches, M., Krupa, M., Teruel, A.: Canard solutions in planar piecewise linear systems with three zones. Dyn. Syst. Int. J. 31, 173–197 (2016). https://doi.org/10.1080/14689367.2015.1079304

    Article  MathSciNet  Google Scholar 

  • Freire, E., Ponce, E., Torres, F.: Canonical discontinuous planar piecewise linear systems. SIAM J. Appl. Dyn. Syst. 11, 181–211 (2012). https://doi.org/10.1137/11083928X

    Article  MathSciNet  Google Scholar 

  • Golubitsky, M., Josic, K., Kaper, T.J.: An unfolding theory approach to bursting in fast-slow systems. In: Broer, H.W., Krauskopf, B., Vegter, G. (eds.) Global Analysis of Dynamical Systems, pp. 277–308. IOP Publishing, Bristol (2001)

    Google Scholar 

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

    Article  MathSciNet  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 Biol. Sci. 221(1222), 87–102 (1984)

    Article  Google Scholar 

  • Izhikevich, E.M.: Neural excitability, spiking and bursting. Int. J. Bifurc. Chaos 10(06), 1171–1266 (2000)

    Article  MathSciNet  Google Scholar 

  • Izhikevich, E.M.: Dynamical Systems in Neuroscience. MIT Press, Cambridge (2007)

    Google Scholar 

  • Konishi, K., Takeuchi, M., Shimizu, T.: Design of external forces for eliminating traveling wave in a piecewise linear Fitzhugh–Nagumo model. Chaos Interdiscip. J. Nonlinear Sci. 21(2), 023101 (2011). https://doi.org/10.1063/1.3545162

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Kuznetsov, Y.A.: Elements of Applied Bifurcation Theory. Applied Mathematical Sciences, 2nd edn. Springer, Berlin (1998)

    Google Scholar 

  • Lecar, H.: Morris–Lecar model. Scholarpedia 2(10), 1333 (2007)

    Article  Google Scholar 

  • Llibre, J., Núñez, E., Teruel, A.: Limit cycles for planar piecewise linear differential systems via first integrals. Qual. Theory Dyn. Syst. 3, 29–50 (2002). https://doi.org/10.1007/BF02969332

    Article  MathSciNet  Google Scholar 

  • Llibre, J., Ponce, E., Teruel, A.E.: Horseshoes near homoclinic orbits for piecewise linear differential systems in \({\mathbb{R} }^{3}\). Int. J. Bifurc. Chaos 17, 1171–1184 (2007). https://doi.org/10.1142/S0218127407017756

    Article  Google Scholar 

  • Mbé, J.H.T., Talla, A.F., Chengui, G.R.G., Coillet, A., Larger, L., Woafo, P., Chembo, Y.K.: Mixed-mode oscillations in slow-fast delayed optoelectronic systems. Phys. Rev. E 91(1), 012902 (2015)

    Article  Google Scholar 

  • Medrado, J.C., Torregrosa, J.: Uniqueness of limit cycles for sewing piecewise linear systems. J. Math. Anal. Appl. 431, 529–544 (2015)

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  • Neishtadt, A.I.: Persistence of stability loss for dynamical bifurcations I. Differ. Equ. 23, 1385–1391 (1987)

    Google Scholar 

  • Neishtadt, A.I.: Persistence of stability loss for dynamical bifurcations II. Differ. Equ. 24, 171–176 (1988)

    Google Scholar 

  • Neishtadt, A.I.: On stability loss delay for dynamical bifurcations. Discrete Contin. Dyn. Syst. Ser. S 2(4), 897 (2009)

    MathSciNet  Google Scholar 

  • Penalva, J., Desroches, M., Teruel, A.E., Vich, C.: Slow passage through a Hopf-like bifurcation in piecewise linear systems: application to elliptic bursting. Chaos Interdiscip. J. Nonlinear Sci. 32(12), 123109 (2022). https://doi.org/10.1063/5.0101778

    Article  MathSciNet  Google Scholar 

  • Perko, L.: Differential Equations and Dynamical Systems. Texts in Applied Mathematics. Springer, New York (1996)

    Book  Google Scholar 

  • Prohens, R., Teruel, A.E., Vich, C.: Slow-fast n-dimensional piecewise linear differential systems. J. Differ. Equ. 260(2), 1865–1892 (2016). https://doi.org/10.1016/j.jde.2015.09.046

    Article  MathSciNet  Google Scholar 

  • Rajesh, S., Ananthakrishna, G.: Relaxation oscillations and negative strain rate sensitivity in the Portevin-Le Chatelier effect. Phys. Rev. E 61(4), 3664 (2000)

    Article  Google Scholar 

  • Rinaldi, S., Muratori, S.: Slow-fast limit cycles in predator–prey models. Ecol. Model. 61(3), 287–308 (1992). https://doi.org/10.1016/0304-3800(92)90023-8

    Article  Google Scholar 

  • Rinzel, J., Ermentrout, G.: Methods in neuronal modeling: from synapses to networks. In: Koch, C., Segev, I. (eds.) Analysis of Neural Excitability and Oscillations, 2nd edn., pp. 251–291. MIT Press, Cambridge (1989)

    Google Scholar 

  • Rinzel, J.: A formal classification of bursting mechanisms in excitable systems. In: International Congress of Mathematicians. Berkeley, California, USA, August 3–11, 1986, vol. II, pp. 1578–1593. American Mathematical Society, Providence (1987)

  • Rinzel, J.: A formal classification of bursting mechanisms in excitable systems. In: Teramoto, E., Yumaguti, M. (eds.) Mathematical Topics in Population Biology, Morphogenesis and Neurosciences. Lecture Notes in Biomathematics, vol. 71, pp. 267–281. Springer, Berlin (1987)

    Chapter  Google Scholar 

  • Rotstein, H.G., Kopell, N., Zhabotinsky, A.M., Epstein, I.R.: Canard phenomenon and localization of oscillations in the Belousov–Zhabotinsky reaction with global feedback. J. Chem. Phys. 119(17), 8824–8832 (2003)

    Article  Google Scholar 

  • Saggio, M.L., Spiegler, A., Bernard, C., Jirsa, V.K.: Fast-slow bursters in the unfolding of a high codimension singularity and the ultra-slow transitions of classes. J. Math. Neurosci. 7(1), 7 (2017)

    Article  MathSciNet  Google Scholar 

  • Shilnikov, A., Kolomiets, M.: Methods of the qualitative theory for the Hindmarsh–Rose model: a case study—a tutorial. Int. J. Bifurc. Chaos 18(08), 2141–2168 (2008)

    Article  MathSciNet  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  Google Scholar 

  • Tonnelier, A., Gerstner, W.: Piecewise linear differential equations and integrate-and-fire neurons: insights from two-dimensional membrane models. Phys. Rev. E: Stat. Nonlinear Soft Matter Phys. 67(2 Pt 1), 021908 (2003). https://doi.org/10.1103/PhysRevE.67.021908

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

Partial financial support was received from the Ministerio de Ciencia, Innovación y Universidades (MCIU) project PID2020-118726GB-I00. AET and CV are also partially supported by the Ministerio de Economia y Competitividad through the project MTM2017-83568-P (AEI/ERDF,EU). JPV is also partially supported by a grant Beques de Mobilitat EDUIB - Santander Universitats.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. Penalva.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Additional information

Communicated by Jeff Moehlis.

Publisher's Note

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

Appendices

Local Flow

In this appendix, we present some local properties of the flow of system (2) when the parameters are contained in the region \({\mathcal {P}}\), see (4).

System (2) can be written in terms of its associated linear system

$$\begin{aligned} \mathbf {{\dot{x}}} = M_{i,j}{\textbf{x}}+b_{i,j}, \end{aligned}$$

where \(M_{i,j}\) and \(b_{i,j}\) are the matrix and the vector components associated to the linear region \(R_{i,j}\). The local flow of (2) in the region \(R_{i,j}\) is determined by the trace, the determinant and the discriminant of \(M_{i,j}\): \(\textrm{tr}_{i,j}\), \(\det _{i,j}\), and \(\Delta _{i,j}\), respectively. Let us denote \({\textbf{p}}_{i,j}=-M_{i,j}^{-1}b_{i,j}\). If \({\textbf{p}}_{i,j}\in R_{i,j}\), this point is an equilibrium point of the system; otherwise, it is a virtual equilibrium (Di Bernardo et al. 2008).

Depending on the region, the matrices, traces, determinants and equilibrium points are as follows. Matrix

$$\begin{aligned} M_{1,1} = \begin{pmatrix} -1/\varepsilon &{} \quad -1/\varepsilon \\ 0 &{}\quad -1 \end{pmatrix}, \end{aligned}$$

with \(\det _{1,1}=1/\varepsilon >0\), \(\textrm{tr}_{1,1}=-1/\varepsilon -1<0\) and \(\Delta _{1,1}=(1/\varepsilon -1)^2\ge 0\). When point \({\textbf{p}}_{1,1}=(I,0)\) is an equilibrium point it is an attractor node. Matrix

$$\begin{aligned} M_{2,1} = \begin{pmatrix} \delta /\varepsilon &{}\quad -1/\varepsilon \\ 0 &{} \quad -1 \end{pmatrix}, \end{aligned}$$

with \(\det _{2,1}=-\delta /\varepsilon >0\), \(\textrm{tr}_{2,1}=\delta /\varepsilon -1<0\) and \(\Delta _{2,1}=(\delta /\varepsilon +1)^2\ge 0\). Again, when point \({\textbf{p}}_{2,1}=(\frac{\lambda +I}{-\delta },0)\) is an equilibrium point it is an attractor node. Matrix

$$\begin{aligned} M_{3,1} = \begin{pmatrix} k/\varepsilon &{}\quad -1/\varepsilon \\ 0 &{}\quad -1 \end{pmatrix}, \end{aligned}$$

with \(\det _{3,1}=-k/\varepsilon <0\). Since the determinant is negative, when the point \({\textbf{p}}_{3,1}=(-\frac{\beta +I}{k},0)\) is an equilibrium, it is a saddle. The eigenvectors associated to the eigenvalues \(\lambda _1=k/\varepsilon \), \(\lambda _2=-1\) are \(v_1=(1,0)\), \(v_2=(k+\varepsilon ,1)\), respectively. These directions over the equilibrium point define the unstable manifold \(W^u({\textbf{p}}_{3,1})\) and the stable manifold \(W^s({\textbf{p}}_{3,1})\), respectively. Matrix

$$\begin{aligned} M_{3,2} = \begin{pmatrix} k/\varepsilon &{}\quad -1/\varepsilon \\ l &{}\quad -1 \end{pmatrix}. \end{aligned}$$

with \(\det _{3,2}=(l-k)/\varepsilon >0\), \(\textrm{tr}_{3,2}=k/\varepsilon -1\) and \(\Delta _{3,2}=\frac{1}{\varepsilon ^2}\Big ((k+\varepsilon )^2-4l\varepsilon \Big )\). When the point \({\textbf{p}}_{3,2}=\frac{1}{k-l}(n-\beta -I,kn-Il-\beta l)\) is an equilibrium point it has either a focus or a node type dynamics. In particular, for sufficiently large l the point \({\textbf{p}}_{3,2}\) is a repelling focus with eigenvalues \(\rho \pm \theta \,\textrm{i}\), where \(\rho =(k-\varepsilon )/(2\varepsilon )\) and \(\theta =\sqrt{4l\varepsilon -(k+\varepsilon )^2}/(2\varepsilon )\). In such a case, the local expression of the orbit of system (2) with initial condition \({\textbf{q}}_0=(x_0,y_0)^T \in R_{3,2}\) is given by:

$$\begin{aligned} \begin{pmatrix} x(t) \\ y(t) \end{pmatrix} = P^{-1}e^{Jt}P({\textbf{q}}_0+A^{-1}{\textbf{a}}_0)-A^{-1}{\textbf{a}}_0, \end{aligned}$$

where

$$\begin{aligned} P^{-1} = \begin{pmatrix} \frac{\rho +1}{l} &{}\quad \frac{\theta }{l} \\ 1 &{}\quad 0 \end{pmatrix}, \quad e^{Jt} = e^{\rho t}\begin{pmatrix} \cos (\theta t) &{}\quad \sin (\theta t) \\ -\sin (\theta t) &{}\quad \cos (\theta t) \end{pmatrix}, \\ A = \begin{pmatrix} k/\varepsilon &{}\quad -1/\varepsilon \\ l &{}\quad -1 \end{pmatrix}, \quad {\textbf{a}}_0 = \left( \frac{1}{\varepsilon }(\beta +I), n\right) ^T. \end{aligned}$$

We finally consider matrix

$$\begin{aligned} M_{4,3} = \begin{pmatrix} -1/\varepsilon &{}\quad -1/\varepsilon \\ 0 &{}\quad -1 \end{pmatrix}, \end{aligned}$$

which satisfies that \(M_{4,3}=M_{1,1}\). Hence, when \({\textbf{p}}_{4,3}=(\gamma -1+I,1)\) is an equilibrium point then it is an attractor node. The expressions of the orbits in the half-space \(\{x>1\}\) are given by

$$\begin{aligned} \begin{aligned}&x(t) = C_1e^{-t/\varepsilon }-\frac{C_2}{1-\varepsilon }e^{-t} - 1 + \gamma + I, \\&y(t) = C_2e^{-t}+1, \end{aligned} \end{aligned}$$
(A1)

where \(C_1,C_2\in {\mathbb {R}}\).

Let us consider the point \({\textbf{p}}=(1,-1+\gamma +I)^T\) on the separation line \(\{x=1\}\) of the region \(R_{4,3}\). Denoting by \(\dot{{\textbf{p}}}\) the vector field of system (2) evaluated at \({\textbf{p}}\) it follows that \(\dot{{\textbf{p}}}=M_{4,3}{\textbf{p}}+b_{4,3}=(0,2-\gamma -I)^T\). Hence, provided \(2-\gamma -I\ne 0\), the set \(\{{\textbf{p}},\dot{{\textbf{p}}}\}\) is a base of \({\mathbb {R}}^2\) called a Krylov base, and the orbits of system (2) can be expressed in terms of this base as \(\Gamma (t)=u(t){\textbf{p}}+v(t)\dot{{\textbf{p}}}\). From Theorem 5 of Llibre et al. (2002), as long as \(\Gamma (t)\subset R_{4,3}\), function

$$\begin{aligned} H(u,v)=(u+\lambda _1v)^{\lambda _2}(u+\lambda _2v)^{-\lambda _1}, \end{aligned}$$
(A2)

is constant along \(\Gamma (t)\), where \(\lambda _1,\lambda _2\) are the eigenvalues of matrix \(M_{4,3}\). Therefore, the point transformation defined by the flow in region \(R_{4,3}\) on the line \(\{x=1\}\) can be expressed as \(\pi (z)\), where the point \({\textbf{p}}+z\dot{{\textbf{p}}}\) maps onto the point \({\textbf{p}}+\pi (z)\dot{{\textbf{p}}}\), through the implicit relation \(H(1,z)=H(1,\pi (z))\).

For a sufficiently smooth function h(z), and to simplify the notation, let us consider the following abuse of notation \(H(z)=H(1,h(z))\). It follows that

$$\begin{aligned} H''(z)&= D (\lambda _1 - \lambda _2) (\lambda _1 h(z) + 1)^{\lambda _2 - 2} (\lambda _2 h(z) + 1)^{-\lambda _1 - 2} \\&\quad \Big (h'(z)^2 (D (\lambda _1 - \lambda _2 + 1) h(z)^2 - 1) \\&\quad - h(z) (\lambda _1 h(z) + 1) (\lambda _2 h(z) + 1) h''(z) \Big ), \end{aligned}$$
$$\begin{aligned} H'''(z)&= D (\lambda _1 h(z) + 1)^{\lambda _2 - 3} (\lambda _2 h(z) + 1)^{-\lambda _1 - 3} \\&\quad \Big ( (\lambda _2 - \lambda _1) h(z) h^{(3)}(z) (\lambda _1 h(z) + 1)^2 (\lambda _2 h(z) + 1)^2 - (\lambda _1 - \lambda _2) h'(z)^3 \\&\quad (D (\lambda _1 - \lambda _2 + 2) h(z) (D (\lambda _1 - \lambda _2 + 1) h(z)^2 - 3) - 2 (T)) \\&\quad + 3 (\lambda _1 - \lambda _2) (\lambda _1 h(z) + 1) (\lambda _2 h(z) + 1) h'(z) h''(z) \\&\quad (D(\lambda _1 - \lambda _2 + 1) h(z)^2 - 1) \Big ), \end{aligned}$$

where \(T=\lambda _1+\lambda _2\) and \(D=\lambda _1\lambda _2\). Hence, by equalling \(H''(0)=H''(\pi (0))\) and \(H'''(0)=H'''(\pi (0))\), we get the first terms of the time series expansion of \(\pi (z)\), where

$$\begin{aligned} \pi (z)= -z + \frac{2T}{3}z^2 + O(z^3), \end{aligned}$$

which describes the flow in \(R_{4,3}\) in a neighbourhood of \(z=0\), corresponding with the coordinate in the Krylov base of the contact point \((1,k+\beta +I)\).

Finally, by direct integration, the solution of system (7) in the half-space \(\{x>1\}\) is given by

$$\begin{aligned} \begin{aligned}&x_{4,3}(t) = C_1e^{-t/\varepsilon }-\frac{C_2}{1-\varepsilon }e^{-t}+\sigma t - 1 + \gamma -\sigma \varepsilon + I_0 \\&y_{4,3}(t) = C_2e^{-t}+1 \\&I(t) = \sigma t + I_0 \end{aligned} \end{aligned}$$
(A3)

We notice that \(I(t) = \sigma t + I_0\) is the same expression for any region. On the other hand, orbits in region \(R_{3,2}\) for the \({\textbf{x}}(t)=(x_{3,2}(t),y_{3,2})^t\) variables are given by

$$\begin{aligned} {\textbf{x}}(t) = P^{-1}e^{Jt}P{\textbf{x}}_0+{\textbf{x}}_p, \end{aligned}$$
(A4)

where P, \(e^{Jt}\) are the ones given previously, \({\textbf{x}}_0\in {\mathbb {R}}^2\) and \({\textbf{x}}_p\) is a particular solution; for instance, in our case we have taken the solution \({\textbf{x}}_p=(a_{p}t+b_{p},(ka_p+\sigma )t+kb_p-\varepsilon a_p+\beta +I_0)\), with \(a_p=\sigma /(l-k)\) and \(b_p=\frac{(k-\varepsilon )\sigma /(l-k)+\sigma +\beta -n+I_0}{l-k}\).

2-Spikes Maximal Canards

In this appendix, we provide all the partial derivatives of the equations in system (34) with respect to the variables \(\{\sigma ,y_1,y_2,y_3,z_1,z_2,z_3,\tau _1,{\tilde{\tau }}_2,\tau _3\}\) and evaluated at the point \((0,{\hat{y}}_1(\varepsilon ),0,{\hat{y}}_1(\varepsilon ),I(\varepsilon ),I(\varepsilon ),I(\varepsilon ),\tau (\varepsilon ),0,\tau (\varepsilon ))\).

We first compute the partial derivative with respect to \(y_1\). Since this variable only appears in the 2nd, 4th and 5th equations, it follows that \(\frac{\textrm{d}\phi _i}{\textrm{d}y_1}=0\) for \(i\ne 2,4,5\). Moreover,

$$\begin{aligned} \frac{\textrm{d}\phi _2}{\textrm{d}y_1}=-1, \quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _4}{\textrm{d}y_1}=\frac{-1}{k+\varepsilon },\quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _5}{\textrm{d}y_1}=0. \end{aligned}$$

Variable \(y_2\) only appears in the 5th, 7th and 8th equations, so

$$\begin{aligned} \frac{\textrm{d}\phi _5}{\textrm{d}y_2}=-1,\quad \frac{\textrm{d}\phi _7}{\textrm{d}y_2}=\frac{e^{ \frac{\tau (\varepsilon )}{\varepsilon } }-e^{-\tau (\varepsilon )}}{1-\varepsilon }, \quad \frac{\textrm{d}\phi _8}{\textrm{d}y_2}=e^{-\tau (\varepsilon )}. \end{aligned}$$

Similarly, variable \(y_3\) only appears in the 8th and 9th equations, so

$$\begin{aligned} \frac{\textrm{d}\phi _8}{\textrm{d}y_3}=-1, \quad \frac{\textrm{d}\phi _9}{\textrm{d}y_3}=\frac{k}{k+\varepsilon }. \end{aligned}$$

Variable \(z_1\) only appears in 1st and 3rd equations, thus

$$\begin{aligned} \frac{\textrm{d}\phi _1}{\textrm{d}z_1}=1-e^{- \frac{\tau (\varepsilon )}{\varepsilon } }, \quad \frac{\textrm{d}\phi _3}{\textrm{d}z_1}=1. \end{aligned}$$

Variable \(z_2\) appears in the 3rd, 4th and 5th equations, thus

$$\begin{aligned} \frac{\textrm{d}\phi _3}{\textrm{d}z_2}=-1, \quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _4}{\textrm{d}z_2}=\frac{1}{k}, \quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _5}{\textrm{d}z_2}=0. \end{aligned}$$

Variable \(z_3\) appears in the 6th, 7th and 9th equations, and so

$$\begin{aligned} \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _6}{\textrm{d}z_3}=0, \quad \frac{\textrm{d}\phi _7}{\textrm{d}z_3}=1-e^{-\frac{\tau (\varepsilon )}{\varepsilon } }, \quad \frac{\textrm{d}\phi _9}{\textrm{d}z_3}=1. \end{aligned}$$

Variable \(\sigma \) appears in the 1st, 3rd, 4th, 6th, 7th and 9th equations, and its derivatives are

$$\begin{aligned}{} & {} \frac{\textrm{d}\phi _1}{\textrm{d}\sigma }=\frac{\textrm{d}\phi _7}{\textrm{d}\sigma }=\varepsilon (e^{- \frac{\tau (\varepsilon )}{\varepsilon } }-1)+\tau (\varepsilon ), \quad \frac{\textrm{d}\phi _3}{\textrm{d}\sigma }=\frac{\textrm{d}\phi _9}{\textrm{d}\sigma }=\tau (\varepsilon ), \\{} & {} \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _4}{\textrm{d}\sigma }=\frac{\varepsilon }{k^2}, \quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _6}{\textrm{d}\sigma }=1. \end{aligned}$$

Variable \(\tau _1\) appears in the 1st, 2nd, 3rd equations, and so

$$\begin{aligned} \frac{\textrm{d}\phi _1}{\textrm{d}\tau _1}=\frac{k+\beta +I(\varepsilon )}{\varepsilon }, \quad \frac{\textrm{d}\phi _2}{\textrm{d}\tau _1}=e^{-\tau (\varepsilon )}, \quad \frac{\textrm{d}\phi _3}{\textrm{d}\tau _1}=0. \end{aligned}$$

Variable \({\tilde{\tau }}_2\) appears in the 4th, 5th and 6th equations, and its derivatives are

$$\begin{aligned} \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _4}{d{\tilde{\tau }}_2}=0,\quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _5}{d{\tilde{\tau }}_2}=0,\quad \lim _{{\tilde{\tau }}_2\rightarrow 0}\frac{\textrm{d}\phi _6}{d{\tilde{\tau }}_2}=z_2-z_3. \end{aligned}$$

Variable \(\tau _3\) only appears in the 7th, 8th and 9th equations, so

$$\begin{aligned} \frac{\textrm{d}\phi _7}{\textrm{d}\tau _3}=\frac{k+\beta +I(\varepsilon )-{\hat{y}}_1(\varepsilon )}{\varepsilon },\quad \frac{\textrm{d}\phi _8}{\textrm{d}\tau _3}=(1-{\hat{y}}_1(\varepsilon ))e^{-\tau (\varepsilon )},\quad \frac{\textrm{d}\phi _9}{\textrm{d}\tau _3}=0. \end{aligned}$$

Homoclinic Connection

In this appendix, we obtain system (32) from system (29), and we show that both systems are equivalent for \(a<1\). Additionally, we compute the Jacobian matrix of system (32) at the solution \({\textbf{v}}_0=(0,\tau (\varepsilon _0),0,0,{\hat{y}}_1(\varepsilon ),I(\varepsilon _0),1,\varepsilon _0)\). Finally, we show that this Jacobian has a minor of maximal range.

In order to do that, let us consider the change of variables (31). From this, we obtain the following expression for the first equation in system (29):

$$\begin{aligned} {\tilde{\phi }}_1 = \frac{1-a}{1-(1-a)k}\left[ (ak+\beta +I) e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1}\Big (\xi _{1,1} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_1)}{\sqrt{1-a}}\right. \\ \left. -\cos (\sqrt{1-a}{\tilde{\tau }}_1)\Big ) - 1 +(k+\beta +I) \right] . \end{aligned}$$

We notice that functions

$$\begin{aligned} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_1)}{\sqrt{1-a}},\quad \cos (\sqrt{1-a}{\tilde{\tau }}_1), \quad e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1}, \end{aligned}$$

are well-defined analytical functions at \(a=1\), and

$$\begin{aligned} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_1)}{\sqrt{1-a}}&= {\tilde{\tau }}_1+\frac{{{{\tilde{\tau }}_1}^{3}} \left( a-1\right) }{6} + O((1-a)^2), \\ \cos (\sqrt{1-a}{\tilde{\tau }}_1)&= 1+{\tilde{\tau }}_1^2(a-1)/2 + O((1-a)^2), \\ e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1}&= 1+\frac{\left( \varepsilon - k\right) }{2\sqrt{\varepsilon }} {\tilde{\tau }}_1 \left( a-1\right) + O((1-a)^2). \end{aligned}$$

Hence, we may consider \({\tilde{\phi }}_1=(1-a){\hat{\phi }}_1\), where \({\hat{\phi }}_1={\tilde{\tau }}_1(k+\beta +I)/\sqrt{\varepsilon }-1+O(1-a)\). This gives us the theoretical value for the variable \({\tilde{\tau }}_1=\sqrt{\varepsilon }/(k+\beta +I)\).

The 2nd equation in system (29) is given by:

$$\begin{aligned} {\tilde{\phi }}_2&= \frac{1-a}{1-(1-a)k}\left[ (ak+\beta +I) \Big (\xi _{2,1} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_1)}{\sqrt{1-a}} e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1} \right. \\&\quad \left. +(1-e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1})\frac{\cos (\sqrt{1-a}{\tilde{\tau }}_1)}{1-a}\Big ) - (1-(1-a)k)g_1 \right] \end{aligned}$$

Again, since the expression \((1-e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_1})\frac{\cos (\sqrt{1-a}{\tilde{\tau }}_1)}{1-a}\) is well defined and analytical at \(a=1\), we can consider \({\tilde{\phi }}_2=(1-a){\hat{\phi }}_2\) where \({\hat{\phi }}_2=\frac{{\tilde{\tau }}_1^2(k+\beta +I)}{2}-g_1+O(1-a)\). This gives us the theoretical value for variable \(g_1=\frac{\varepsilon }{2(k+\beta +I)}\).

The 3rd and 4th equations in system (29) become

$$\begin{aligned} {\hat{\phi }}_3&= (2-\gamma -I)(e^{- \frac{\tau _2}{\varepsilon } }-1)+\frac{1-(1-a)g_1}{1-\varepsilon } (e^{-\tau _2}-e^{- \frac{\tau _2}{\varepsilon } }) = 0 \\ {\hat{\phi }}_4&= -(1-(1-a)g_1)e^{-\tau _2}+1 - {\hat{y}}_1(\varepsilon ) - (1-a)g_2 = 0. \end{aligned}$$

The 5th equation is then given by:

$$\begin{aligned} {\tilde{\phi }}_5 = \frac{1-a}{1-(1-a)k}\left[ (ak+\beta +I) e^{\frac{ \rho }{\theta } \sqrt{1-a}{\tilde{\tau }}_3}\Big (\xi _{5,1} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_3)}{\sqrt{1-a}}\right. \\ \left. -\cos (\sqrt{1-a}{\tilde{\tau }}_3)\Big ) - 1 +(k+\beta +I) \right] , \end{aligned}$$

which can be simplified again as \({\tilde{\phi }}_5=(1-a){\hat{\phi }}_5\) where \({\hat{\phi }}_5=-(k+\beta +I)\sqrt{\varepsilon }{\tilde{\tau }}_3/k-1+O(1-a)\). This gives us the value \({\tilde{\tau }}_3=-k/(\sqrt{\varepsilon }(k+\beta +I))\). Finally, \({\tilde{\phi }}_6\) is given by

$$\begin{aligned} {\tilde{\phi }}_6&= \frac{1-a}{1-(1-a)k}\left[ (ak+\beta +I) e^{\frac{\rho }{\theta }\sqrt{1-a}{\tilde{\tau }}_3}\Big ( \xi _{6,1} \frac{\sin (\sqrt{1-a}{\tilde{\tau }}_3)}{\sqrt{1-a}} \right. \\&\quad + \xi _{6,2} \cos (\sqrt{1-a}{\tilde{\tau }}_3)\Big ) + \frac{1}{1-a}((ak+\beta +I) \\&\quad - (1-(1-a)k)({\hat{y}}_1(\varepsilon )+(1-a)g_2)) \Big ]. \end{aligned}$$

This expression can be written as \({\tilde{\phi }}_6=(1-a){\hat{\phi }}_6\) where \({\hat{\phi }}_6=-(k+\beta +I){\tilde{\tau }}_3\sqrt{\varepsilon }(\sqrt{\varepsilon }{\tilde{\tau }}_3+2(k+\varepsilon ))/(2k)-g_2+O(1-a)\). This provides the value \(g_2=(2(k+\varepsilon )-k/(k+\beta +I))/2\).

We conclude that system (32) is equivalent to system (29) when \(a<1\). Moreover, system (32) is defined at \(a=1\) and

$$\begin{aligned} \tilde{{\textbf{v}}}_0=\left( \frac{\sqrt{\varepsilon }}{k+\beta +I}, \tau , -\frac{k/\sqrt{\varepsilon }}{k+\beta +I}, \frac{\varepsilon /2}{k+\beta +I}, k+\varepsilon -\frac{k/2}{k+\beta +I}, I,1,\varepsilon \right) , \end{aligned}$$

is a solution, where \(\tau =\tau (\varepsilon )\) and \(I=I(\varepsilon )\) are given in Lemma 3.

Let us now compute the Jacobian matrix of system (32) at \(\tilde{{\textbf{v}}}_0\) and show that there exists a minor with maximal range. In particular, the partial derivatives of equation \({\hat{\phi }}_i\) with \(i=1,2,\ldots ,6\) with respect to the variables \(({\tilde{\tau }}_1,\tau _2,{\tilde{\tau }}_3,g_1,g_2,I)\) produce the minor

$$\begin{aligned} A= \begin{pmatrix} \frac{k+\beta +I}{\sqrt{\varepsilon }} &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad I_{1} \\ {\tilde{\tau }}_{1,2} &{} \quad 0 &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad I_{2} \\ 0 &{} \quad \frac{k+\beta +I-1+e^{-\tau }}{\varepsilon } &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad 1-e^{-\frac{\tau }{\varepsilon }} \\ 0 &{} \quad e^{-\tau } &{} \quad 0 &{} \quad 0 &{} \quad 0 &{} \quad -\frac{k+\varepsilon }{k} \\ 0 &{} \quad 0 &{} \quad \frac{(k+\beta +I)\sqrt{\varepsilon }}{k} &{} \quad 0 &{} \quad 0 &{} \quad I_{5} \\ 0 &{} \quad 0 &{} \quad {\tilde{\tau }}_{3,6} &{} \quad 0 &{} \quad -1 &{} \quad I_{6} \\ \end{pmatrix} \end{aligned}$$

where not all the entries has been explicitly computed. In fact, \({\tilde{\tau }}_{i,j}\) stands for \(\frac{\partial {\hat{\phi }}_j}{\partial {\tilde{\tau }}_i}\) and \(I_j\) stands for \(\frac{\partial {\hat{\phi }}_j}{\partial I}\). Therefore,

$$\begin{aligned} \det (A)= -\frac{(k+\beta +I)^2}{k} \begin{vmatrix} \dfrac{k+\beta +I-1+e^{-\tau }}{\varepsilon }&1-e^{- \frac{\tau }{\varepsilon } } \\ e^{-\tau }&-\dfrac{k+\varepsilon }{k} \end{vmatrix} \end{aligned}$$
(C5)

where \((k+\beta +I)\ne 0\) and the second term is precisely the determinant calculated in Lemma 3, which is nonzero.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Penalva, J., Desroches, M., Teruel, A.E. et al. Dynamics of a Piecewise-Linear Morris–Lecar Model: Bifurcations and Spike Adding. J Nonlinear Sci 34, 52 (2024). https://doi.org/10.1007/s00332-024-10029-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-024-10029-3

Keywords

Mathematics Subject Classification

Navigation