Skip to main content
Log in

Multiscale Geometry of the Olsen Model and Non-classical Relaxation Oscillations

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

Abstract

We study the Olsen model for the peroxidase–oxidase reaction. The dynamics is analyzed using a geometric decomposition based on multiple timescales. The Olsen model is four-dimensional, not in a standard form required by geometric singular perturbation theory and contains multiple small parameters. These three obstacles are the main challenges we resolve by our analysis. Scaling and the blow-up method are used to identify several subsystems. The results presented here provide a rigorous analysis for two oscillatory modes. In particular, we prove the existence of non-classical relaxation oscillations in two cases. The analysis is based on desingularization of lines of transcritical and submanifolds of fold singularities in combination with an integrable relaxation phase. In this context, our analysis also explains an assumption that has been utilized, based purely on numerical reasoning, in a previous bifurcation analysis by Desroches et al. (Discret Contin Dyn Syst S 2(4):807–827, 2009). Furthermore, the geometric decomposition we develop forms the basis to prove the existence of mixed-mode and chaotic oscillations in the Olsen model, which will be discussed in more detail in future work.

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

Similar content being viewed by others

Notes

  1. Alexandra Milik (1964–2014) studied the Olsen model in her PhD thesis under the supervision of the second author. She discovered the important role of canard solutions for mixed-mode oscillations. However, at that time, the tools needed for the analysis of the occurring complicated partially non-hyperbolic slow manifolds governing the essential dynamics had not been sufficiently well developed for a complete analysis.

References

  • Aguda, B.D., Larter, R.: Periodic-chaotic sequences in a detailed mechanism of the peroxidase-oxidase reaction. J. Am. Chem. Soc. 113, 7913–7916 (1991)

    Article  Google Scholar 

  • Aguda, B.D., Larter, R., Clarke, B.L.: Dynamic elements of mixed-mode oscillations and chaos in a peroxidase-oxidase network. J. Chem. Phys. 90(8), 4168–4175 (1989)

    Article  Google Scholar 

  • Arnold, V.I.: Encyclopedia of Mathematical Sciences: Dynamical Systems V. Springer, Berlin (1994)

    Book  Google Scholar 

  • Baer, S.M., Erneux, T.: Singular Hopf bifurcation to relaxation oscillations I. SIAM J. Appl. Math. 46(5), 721–739 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  • Barkley, D.: Slow manifolds and mixed-mode oscillations in the Belousov–Zhabotinskii reaction. J. Chem. Phys. 89(9), 5547–5559 (1988)

    Article  MathSciNet  Google Scholar 

  • Benoît, E.: Canards et enlacements. Publ. Math. IHES 72, 63–91 (1990)

    Article  MATH  Google Scholar 

  • Benoît, E., Callot, J.L., Diener, F., Diener, M.: Chasse au canards. Collect. Math. 31, 37–119 (1981)

    Google Scholar 

  • Broer, H.W., Kaper, T.J., Krupa, M.: Geometric desingularization of a cusp singularity in slow–fast systems with applications to Zeeman’s examples. J. Dyn. Differ. Equ. 25(4), 925–958 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Bronnikova, T.V., Fed’kina, V.R., Schaffer, W.M., Olsen, L.F.: Period-doubling bifurcations and chaos in a detailed model of the peroxidase-oxidase reaction. J. Phys. Chem. 99(23), 9309–9312 (1995)

    Article  Google Scholar 

  • Bronnikova, T.V., Schaffer, W.M., Olsen, L.F.: Nonlinear dynamics of the peroxidase-oxidase reaction. I. Bistability and bursting oscillations at low enzyme concentrations. J. Phys. Chem. B 105, 310–321 (2001)

    Article  Google Scholar 

  • Cartwright, M.L., Littlewood, J.E.: On non-linear differential equations of second order. I. The equation \(\ddot{y}-k(1-y^2)\dot{y}+y=b\lambda k\cos (\lambda t+a), k\) large. J. Lond. Math. Soc. 20, 180–189 (1945)

    Article  MathSciNet  MATH  Google Scholar 

  • Cartwright, M.L., Littlewood, J.E.: On non-linear differential equations of second order. II. The equation \(\ddot{y}-kf(y,\dot{y})+g(y, k)=p(t), k{\>}0, f(y) \ge 1\). Ann. Math. 48(2), 472–494 (1947)

    Article  MathSciNet  MATH  Google Scholar 

  • Chiba, H.: Periodic orbits and chaos in fast–slow systems with Bogdanov–Takens type fold points. J. Differ. Equ. 250, 112–160 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Degn, H., Olsen, L.F., Perram, J.W.: Bistability, oscillation, and chaos in an enzyme reaction. Ann. NY Acad. Sci 316(1), 623–637 (1979)

    Article  Google Scholar 

  • Desroches, M., Guckenheimer, J., Kuehn, C., Krauskopf, B., Osinga, H., Wechselberger, M.: Mixed-mode oscillations with multiple time scales. SIAM Rev. 54(2), 211–288 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Desroches, M., Krauskopf, B., Osinga, H.M.: The geometry of mixed-mode oscillations in the Olsen model for the peroxidase-oxidase reaction. DCDS-S 2(4), 807–827 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • di Bernardo, M., Budd, C.J., Champneys, A.R., Kowalczyk, P.: Piecewise-Smooth Dynamical Systems, volume 163 of Applied Mathematical Sciences. Springer, Berlin (2008)

    Google Scholar 

  • Diener, M.: The canard unchained or how fast/slow dynamical systems bifurcate. Math. Intell. 6, 38–48 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  • Dumortier, F.: Techniques in the theory of local bifurcations: blow-up, normal forms, nilpotent bifurcations, singular perturbations. In: Schlomiuk, D. (ed.) Bifurcations and Periodic Orbits of Vector Fields, pp. 19–73. Kluwer, Dortrecht (1993)

    Chapter  Google Scholar 

  • Dumortier, F., Roussarie, R.: Canard cycles and center manifolds, vol. 121 of Memoirs American Mathematical Society AMS (1996)

  • Eckhaus, W.: Relaxation oscillations including a standard chase on French ducks. Lect. Notes Math. 985, 449–494 (1983)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Geest, T., Olsen, L.F., Steinmetz, C.G., Larter, R., Schaffer, : Nonlinear analysis of periodic and chaotic time series from the peroxidase-oxidase reaction. J. Phys. Chem. 97, 8431–8441 (1993)

    Article  Google Scholar 

  • Geest, T., Steinmetz, C.G., Larter, R., Olsen, L.F.: Period-doubling bifurcations and chaos in an enzyme reaction. J. Phys. Chem. 96, 5678–5680 (1992)

    Article  Google Scholar 

  • Grasman, J.: Asymptotic Methods for Relaxation Oscillations and Applications. Springer, Berlin (1987)

    Book  MATH  Google Scholar 

  • Guckenheimer, J., Holmes, P.: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer, New York (1983)

    Book  MATH  Google Scholar 

  • Guckenheimer, J., Scheper, C.: A geometric model for mixed-mode oscillations in a chemical system. SIAM J. Appl. Dyn. Syst. 10(1), 92–128 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Guckenheimer, J., Wechselberger, M., Young, L.-S.: Chaotic attractors of relaxation oscillations. Nonlinearity 19, 701–720 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  • Gucwa, I., Szmolyan, P.: Geometric singular perturbation analysis of an autocatalator model. DCDS-S 2(4), 783–806 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Haiduc, R.: Horseshoes in the forced van der Pol system. Nonlinearity 22, 213–237 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Harvey, E., Kirk, V., Osinga, H.M., Sneyd, J., Wechselberger, M.: Understanding anomalous delays in a model of intracellular calcium dynamics. Chaos 20, 045104 (2010)

    Article  MathSciNet  Google Scholar 

  • Hauck, T., Schneider, F.W.: Mixed-mode and quasiperiodic oscillations in the peroxidase-oxidase reaction. J. Phys. Chem. 97, 391–397 (1993)

    Article  Google Scholar 

  • Hauck, T., Schneider, F.W.: Chaos in a Farey sequence through period doubling in the peroxidase-oxidase reaction. J. Phys. Chem. 98, 2072–2077 (1994)

    Article  Google Scholar 

  • Hauser, M.J.B., Olsen, L.F.: Mixed-mode oscillations and homoclinic chaos in an enzyme reaction. J. Chem. Soc. Faraday Trans. 92(16), 2857–2863 (1996)

    Article  Google Scholar 

  • Hauser, M.J.B., Olsen, L.F., Bronnikova, T.V., Schaffer, W.M.: Routes to chaos in the peroxidase-oxidase reaction: period-doubling and period-adding. J. Phys. Chem. B 101, 5075–5083 (1997)

    Article  Google Scholar 

  • Hirsch, M.W., Pugh, C.C., Shub, M.: Invariant Manifolds. Springer, Berlin (1977)

    MATH  Google Scholar 

  • Huber, A., Szmolyan, P.: Geometric singular perturbation analysis of the Yamada model. SIAM J. Appl. Dyn. Syst. 4(3), 607–648 (2005)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Jones, C.K.R.T.: Geometric singular perturbation theory. In: Dynamical Systems (Montecatini Terme, 1994), volume 1609 of Lecture Notes Mathematics, pp. 44–118. Springer, Berlin (1995)

  • Kaper, T.J.: An introduction to geometric methods and dynamical systems theory for singular perturbation problems. Analyzing multiscale phenomena using singular perturbation methods. In: Cronin, J., O’Malley, R.E. (eds.) Analyzing Multiscale Phenomena Using Singular Perturbation Methods, pp. 85–131. Springer, UK (1999)

    Chapter  Google Scholar 

  • Kaper, T.J., Jones, C.K.R.T.: A primer on the exchange lemma for fast–slow systems. In: Multiple-Time-Scale Dynamical Systems, pp. 65–88. Springer, Berlin (2001)

  • Kosiuk, I., Szmolyan, P.: Scaling in singular perturbation problems: blowing-up a relaxation oscillator. SIAM J. Appl. Dyn. Syst. 10(4), 1307–1343 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Kosiuk, I., Szmolyan, P.: A new type of relaxation oscillations in a model of the mitotic oscillator (preprint) (2014)

  • Krupa, M., Popovic, N., Kopell, N.: Mixed-mode oscillations in three time-scale systems: a prototypical example. SIAM J. Appl. Dyn. Syst. 7(2), 361–420 (2008)

    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.: Extending slow manifolds near transcritical and pitchfork singularities. Nonlinearity 14, 1473–1491 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa, M., Szmolyan, P.: Geometric analysis of the singularly perturbed planar fold. In: Jones, C.K.R.T., Khibnik, A.I. (eds.) Multiple-Time-Scale Dynamical Systems, The IMA Volumes in Mathematics and its Applications, vol. 122, pp. 89–116. Springer, New York (2001)

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

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa, M., Vidal, A., Desroches, M., Clément, F.: Mixed-mode oscillations in a multiple time scale phantom bursting system. SIAM J. Appl. Dyn. Syst. 11(4), 1458–1498 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Kuehn, C.: On decomposing mixed-mode oscillations and their return maps. Chaos 21(3), 033107 (2011)

    Article  Google Scholar 

  • Kuehn, C.: Normal hyperbolicity and unbounded critical manifolds. Nonlinearity 27(6), 1351–1366 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Larter, R., Bush, C.L., Lonis, T.R., Aguda, B.D.: Multiple steady states, complex oscillations, and the devil’s staircase in the peroxidase-oxidase reaction. J. Chem. Phys. 87(10), 5765–5771 (1987)

    Article  Google Scholar 

  • Larter, R., Hemkin, S.: Further refinements of the peroxidase-oxidase oscillator mechanism: Mixed-mode oscillations and chaos. J. Phys. Chem. 100, 18924–18930 (1996)

    Article  Google Scholar 

  • Larter, R., Steinmetz, C.G.: Chaos via mixed-mode oscillations. Phil. Trans. R. Soc. Lond. A 337, 291–298 (1991)

    Article  MATH  Google Scholar 

  • Liu, D.: Exchange lemmas for singular perturbation problems with certain turning points. J. Differ. Equ. 167, 134–180 (2000)

    Article  MATH  Google Scholar 

  • Milik, A.: Mixed-mode oscillations in chemical systems. PhD thesis, Vienna University of Technology, Vienna, Austria (1998)

  • Mishchenko, E.F., Kolesov, YuS, Kolesov, AYu., Rozov, NKh: Asymptotic Methods in Singularly Perturbed Systems. Plenum Press, Berlin (1994)

    Book  MATH  Google Scholar 

  • Mishchenko, E.F., Rozov, NKh: Differential Equations with Small Parameters and Relaxation Oscillations (translated from Russian). Plenum Press, Berlin (1980)

    Book  Google Scholar 

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

    MathSciNet  Google Scholar 

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

    MathSciNet  Google Scholar 

  • Olsen, L.F.: An enzyme reaction with a strange attractor. Phys. Lett. A 94(9), 454–457 (1983)

    Article  Google Scholar 

  • Olsen, L.F., Degn, H.: Oscillatory kinetics of the peroxidase-oxidase reaction in an open system. Experimental and theoretical studies. Biochim. Biophys. Acta 523(2), 321–334 (1978)

    Article  Google Scholar 

  • Olson, D.L., Williksen, E.P., Scheeline, A.: An experimentally based model of the peroxidase-NADH biochemical oscillator: an enzyme-mediated chemical switch. J. Am. Chem. Soc. 117, 2–15 (1995)

    Article  Google Scholar 

  • Rinzel, J.: A Formal Classification of Bursting Mechanisms in Excitable Systems. Proc. Int. Congress Math, Berkeley (1986)

    Google Scholar 

  • Schaffer, W.M., Bronnikova, T.V., Olsen, L.F.: Nonlinear dynamics of the peroxidase-oxidase reaction. II. Compatibility of an extended model with previously reported model-data correspondences. J. Phys. Chem. 105, 5331–5340 (2001)

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Scheeline, A., Olson, D.L., Williksen, E.P., Horras, G.A., Klein, M.L., Larter, R.: The peroxidase-oxidase oscillator and its constituent chemistries. Chem. Rev. 97, 739–756 (1997)

    Article  Google Scholar 

  • Steinmetz, C.G., Geest, T., Larter, R.: Universality in the peroxidase-oxidase reaction: period doublings, chaos, period three, and unstable limit cycles. J. Phys. Chem. 97, 5649–5653 (1993)

    Article  Google Scholar 

  • Steinmetz, C.G., Larter, R.: The quasiperiodic route to chaos in a model of the peroxidase-oxidase reaction. J. Phys. Chem. 94(2), 1388–1396 (1991)

    Article  MathSciNet  Google Scholar 

  • Szmolyan, P., Wechselberger, M.: Canards in \({\mathbb{R}}^3\). J. Differ. Equ. 177, 419–453 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Thompson, D.R., Larter, R.: Multiple time scale analysis of two models for the peroxidase-oxidase reaction. Chaos 5(2), 448–457 (1995)

    Article  Google Scholar 

  • Tikhonov, A.N.: Systems of differential equations containing small parameters in the derivatives. Mat. Sbornik N. S. 31, 575–586 (1952)

    MathSciNet  Google Scholar 

  • van der Pol, B.: A theory of the amplitude of free and forced triode vibrations. Radio Rev. 1, 701–710 (1920)

    Google Scholar 

  • van der Pol, B.: On relaxation oscillations. Philos. Mag. 7, 978–992 (1926)

    Google Scholar 

  • Wechselberger, M.: A propos de canards (apropos canards). Trans. Am. Math. Soc. 364, 3289–3309 (2012)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

CK would like to thank the Austrian Academy of Sciences (ÖAW) for support via an APART fellowship. CK and PS would like to thank the European Commission (EC/REA) for support by a Marie-Curie International Re-integration Grant.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Kuehn.

Additional information

Communicated by Sue Ann Campbell.

Dedicated to the memory of Alexandra Milik.

Appendices

Appendix 1: Normally Hyperbolicity and Fast–Slow Systems

We only recall the basic definitions and results about fast–slow systems. There are several standard references that detail many parts of the theory (Jones 1995; Kaper and Jones 2001; Mishchenko and Rozov 1980; Desroches et al. 2012; Arnold 1994). A fast–slow system of ordinary differential equations (ODEs) is given by:

$$\begin{aligned} \begin{array}{rcrcl} \epsilon \dot{x}&{}=&{}\epsilon \frac{\mathrm{d}x}{\mathrm{d}\tau }&{}=&{}f(x,y,\epsilon ),\\ \dot{y}&{}=&{}\frac{\mathrm{d}y}{\mathrm{d}\tau }&{}=&{}g(x,y,\epsilon ),\\ \end{array} \end{aligned}$$
(76)

where \(x\in {\mathbb {R}}^m\) are fast variables, \(y\in {\mathbb {R}}^n\) are slow variables, and \(0<\epsilon \ll 1\) is a small parameter representing the ratio of timescales. The maps \(f,g\) are assumed to be sufficiently smooth. Equation (76) can be rewritten by changing from the slow timescale \(\tau \) to the fast timescale \(t=\tau /\epsilon \)

$$\begin{aligned} \begin{aligned} x'&=\textstyle \frac{\mathrm{d}x}{\mathrm{d}t}=f(x,y,\epsilon ),\\ y'&=\textstyle \frac{\mathrm{d}y}{\mathrm{d}t}=\epsilon ~g(x,y,\epsilon ).\\ \end{aligned} \end{aligned}$$
(77)

The singular limit \(\epsilon \rightarrow 0\) of (77) yields the fast subsystem ODEs parametrized by the slow variables \(y\). Setting \(\epsilon \rightarrow 0\) in (76) gives a differential–algebraic equation (DAE), called the slow subsystem, on the critical manifold \({\mathcal {C}}_0:=\{(x,y)\in {\mathbb {R}}^{m+n}:f(x,y,\epsilon )=0\}\). Concatenations of fast and slow subsystem trajectories are called candidates (Benoît 1990; Haiduc 2009).

A subset \({\mathcal {S}}\subset {\mathcal {C}}\) is called normally hyperbolic if the \(m\times m\) total derivative matrix \((D_xf)(p)\) is hyperbolic for \(p\in {\mathcal {S}}\). A normally hyperbolic subset \({\mathcal {S}}\) is attracting if all eigenvalues of \((D_xf)(p)\) have negative real parts for \(p\in {\mathcal {S}}, {\mathcal {S}}\) is called repelling if all eigenvalues have positive real parts, and \({\mathcal {S}}\) is of saddle type if there are positive and negative eigenvalues. On normally hyperbolic parts of \({\mathcal {C}}\), the implicit function theorem applies to \(f(x,y,0)=0\) providing a map \(h_0(y)=x\) so that \({\mathcal {C}}\) can be expressed (locally) as a graph.

Theorem 9.1

(Fenichel’s Theorem Fenichel (1979); Jones (1995); Tikhonov (1952) Suppose \({\mathcal {S}}={\mathcal {S}}_0\) is a compact normally hyperbolic submanifold (possibly with boundary) of the critical manifold \({\mathcal {C}}_0\). Then, for \(\epsilon >0\) sufficiently small, there exists a locally invariant manifold \({\mathcal {S}}_\epsilon \) diffeomorphic to \({\mathcal {S}}_0\). \({\mathcal {S}}_\epsilon \) has a distance of \({\mathcal {O}}(\epsilon )\) from \({\mathcal {S}}_0\) and the flow on \({\mathcal {S}}_\epsilon \) converges to the slow flow as \(\epsilon \rightarrow 0\).

The distance between \({\mathcal {S}}_\epsilon \) and \({\mathcal {S}}_0\) can be expressed in the Hausdorff metric or a suitable \(C^r\)-norm (using the map \(h_0\) and its perturbation \(h_\epsilon \)). A manifold \({\mathcal {S}}_\epsilon \) provided by Fenichel’s Theorem is called a slow manifold. Slow manifolds are usually not unique, but different slow manifolds lie at a distance \({\mathcal {O}}(e^{-K/\epsilon })\) for some constant \(K>0\). Often we shall make a choice of compact subset and choice of slow manifold without further notice, indicating that the choice does not matter for the asymptotic analysis performed.

A trajectory is called a maximal canard if it lies in the intersection of an attracting and a repelling slow manifold. Canards were first investigated by a group of French mathematicians (Benoît et al. 1981) using nonstandard analysis. Later also asymptotic (Eckhaus 1983; Baer and Erneux 1986) and geometric (Dumortier and Roussarie 1996; Krupa and Szmolyan 2001) methods have been developed to understand canard orbits.

Appendix 2: Geometric Desingularization

Here we shall briefly review the basic strategy for the blow-up approach for geometric desingularization of fast–slow systems. Details on the classical, single-scale, method can be found, e.g., in Dumortier (1993). The classical blow-up was first introduced into fast–slow systems in Dumortier and Roussarie (1996). Further developments can be found in Krupa and Szmolyan (2001), see also the introduction in Krupa and Szmolyan (2001).

The starting point is to write the system (77) as follows

$$\begin{aligned} \begin{aligned} x'&=f(x,y,\epsilon ),\\ y'&=\epsilon ~ g(x,y,\epsilon ),\\ \epsilon '&=0.\\ \end{aligned} \end{aligned}$$
(78)

Let us denote the vector field defined by (78) as \(X\), i.e., \(X\) is a mapping

$$\begin{aligned} X:{\mathbb {R}}^{m+n}\times [0,\epsilon _0)\rightarrow T\left( {\mathbb {R}}^{m+n}\times [0,\epsilon _0)\right) \end{aligned}$$

where \(T(\cdot )\) indicates the tangent bundle. Further equations for parameters could be appended to (78) as well, if necessary. Suppose (78) has an equilibrium point for \(\epsilon =0\), or more generally a submanifold \({\mathcal {M}}=\{f=0\}\) of equilibria in \({\mathbb {R}}^{m+n}\times \{\epsilon =0\}\). If \((D_xf)(p)\) is not a hyperbolic matrix for each \(p\in {\mathcal {M}}\), the equilibrium (manifold) \({\mathcal {M}}\) is degenerate and classical linearization results do not apply directly to (78).

The blow-up technique is based on replacing \({\mathcal {M}}\) by a, usually more complicated, manifold \(\bar{{\mathcal {M}}}\) and using a map

$$\begin{aligned} \Phi :\bar{{\mathcal {M}}}\rightarrow {\mathcal {M}}\end{aligned}$$

which induces a vector field \(\bar{X}\) on \(\bar{{\mathcal {M}}}\) via the pushforward \(\Phi _*\) and the condition \(\Phi _*(\bar{X})=X\). Using a good choice for \(\bar{{\mathcal {M}}}\), one may often analyze the blown-up vector field \(\bar{X}\) as it is possible that invariant manifolds of \(\bar{X}\) are now (partially) hyperbolic.

As an example, consider the classical case when \((x,y)\in {\mathbb {R}}^2\) and \(f(x,y)=y-x^2\) is the (truncated) normal form of a fold bifurcation. The origin \((x,y,\epsilon )=(0,0,0)\) is the important non-hyperbolic point, and the standard choice is to use a sphere for geometric desingularization \(\bar{{\mathcal {M}}}:=S^2\times [0,r_0)\) for some constant \(r_0>0\) or \(r_0=+\infty \). Therefore, one has essentially inserted a sphere at the origin, see also Dumortier and Roussarie (1996), Krupa and Szmolyan (2001).

Although one could try to find a suitable global parametrization of \(\bar{{\mathcal {M}}}\), this is usually not very convenient for calculations. Instead, one uses charts \(\kappa _j:\bar{{\mathcal {M}}}\rightarrow {\mathbb {R}}^{m+n+1}\) of \(\bar{{\mathcal {M}}}\) for the calculations, which is illustrated by the following important diagram

which commutes. Hence, one may just try to calculate the map \(\varphi _j\) and obtain a vector field on \({\mathbb {R}}^{m+n+1}\) by applying the coordinate change

$$\begin{aligned} (x_j,y_j,\epsilon _j)=\varphi ^{-1}(x,y,\epsilon ). \end{aligned}$$

One may often, via a good choice of \(\bar{{\mathcal {M}}}\) and chart maps \(\kappa _j\), compute the vector fields in \((x_j,y_j,\epsilon _j)\)-coordinates. The same remark applies to the transition maps between different charts \(\kappa _{jk}\). Section 3 carries out these calculations for a submanifold of fold points in the Olsen model.

Appendix 3: An Auxiliary Center Manifold Reduction

Here we present the details for the center manifold calculation for (27). We drop the sub- and superscripts of \((r_1,y_1,\epsilon _1)\) and \((\alpha _1^*,b_1^*)\) for notational convenience; all variables and constants used in this section are temporary and should not be confused with notation within the main manuscript. Reordering the variables and translating (27) via \(Y=y-1/(3ab)\) yields

$$\begin{aligned} \begin{aligned} r'&= r\left[ \epsilon (b-\xi )+3abY\right] =:f_1(r,\epsilon ,Y),\\ \epsilon '&=-\epsilon \left[ \epsilon (b-\xi )+3abY\right] =:f_2(r,\epsilon ,Y)\\ Y'&= f_3(r,\epsilon ,Y),\\ \end{aligned} \end{aligned}$$
(79)

where the function \(f_3\) is given by

$$\begin{aligned} f_3(r,\epsilon ,Y):=\kappa \epsilon (1-[Y+1/(3ab)][1+a^*_1b^*_1]) -2[Y+1/(3ab)]\left( \epsilon (b-\xi )+3aby\right) . \end{aligned}$$

Let \(z:=(r,\epsilon ,Y)^T\) and consider

$$\begin{aligned} A:=\left. D_z(z')\right| _{(0,0,0)}= \left( \begin{array}{ccc} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 \\ 0 &{} K &{} -2 \\ \end{array} \right) \quad \text {and}\quad M:=\left( \begin{array}{ccc} 1 &{} 0 &{} 0 \\ 0 &{} -2/K &{} 0 \\ 0 &{} 1 &{} 1 \\ \end{array} \right) , \end{aligned}$$

where \(K:=-(2 b + \kappa - 2 a b \kappa - 2 \xi )/(3 a b)\). Let \((x_1,x_2,\tilde{y})^T=\tilde{z}=M^{-1}z\) and observe that \(M^{-1}AM=J\in {\mathbb {R}}^{3\times 3}\) with \(J_{33}=-2\) and \(J_{ij}=0\) otherwise. Set \(\tilde{z}=(x_1,x_2,\tilde{y})=M^{-1}z\) so that

$$\begin{aligned} \begin{array}{ccccc} \left( \begin{array}{c}x_1'\\ x_2'\\ \end{array}\right) &{}=&{} \left( \begin{array}{cc}0 &{} 0\\ 0&{} 0\\ \end{array}\right) \left( \begin{array}{c}x_1\\ x_2\\ \end{array}\right) &{}+&{} \left( \begin{array}{c}F_1(x_1,x_2,\tilde{y})\\ F_2(x_1,x_2,\tilde{y})\\ \end{array}\right) \\ \tilde{y}'&{}=&{} -2\tilde{y}&{}+&{} G(x_1,x_2,\tilde{y}),\\ \end{array} \end{aligned}$$
(80)

where \((F_{1},F_2,G)^T=(0,0,2\tilde{y})^T+M^{-1}(f_1(M\tilde{z}), f_2(M\tilde{z}),f_3(M\tilde{z}))^T\). The system (80) is in the standard form for center manifold theory (Guckenheimer and Holmes 1983). The usual perturbation ansatz is \(\tilde{y}=h(x_1,x_2)=k_{11}x_1^2+k_{12}x_1x_2+k_{22}x_2^2+{\mathcal {O}}(3),\) where \({\mathcal {O}}(3):={\mathcal {O}}(x_1^3,x_1^2x_2,x_1x_2^2,x_2^3)\). The defining invariance equation for the center manifold with \(x=(x_1,x_2)^T\) and \(F=(F_1,F_2)^T\) is

$$\begin{aligned} Dh(x)F(x,h(x))=-2h(x)+G(x,h(x)) \end{aligned}$$
(81)

since the \(x'\)-equations in (80) have no linear term. Collecting terms of order \({\mathcal {O}}(x_1^2)\) in (81) gives \(k_{11}=0\) and the \({\mathcal {O}}(x_1x_2)\)-terms give \(k_{12}=0\). For \({\mathcal {O}}(x_2^2)\) equation (81) and \(k_{11}=0=k_{12}\) imply

$$\begin{aligned} k_{22}=\frac{3ab(1+4ab)\kappa }{4(b-\xi )+2\kappa (1-2ab)}. \end{aligned}$$

Transforming back to the variables \((r,\epsilon ,y)\) via the matrix \(M\) and translation yields the center manifold

$$\begin{aligned} y=\frac{1}{3ab}+\epsilon \frac{2(\xi -b)+\kappa (2ab-1)}{6ab}+k_{22}\frac{K^2}{4}\epsilon ^2+{\mathcal {O}}(3), \end{aligned}$$

where \({\mathcal {O}}(3)={\mathcal {O}}(r^3,r^2\epsilon ,r\epsilon ^2,\epsilon ^3)\). Computing

$$\begin{aligned} k_{22}\frac{K^2}{4}=\frac{\kappa (1+4ab)}{24ab}(2(b-\xi )+\kappa (1-2ab)) \end{aligned}$$

yields Proposition 3.9.

Appendix 4: Another Auxiliary Center Manifold Reduction

As for the center manifold reduction in Appendix 3, we present some of the important details for the center manifold calculation. As before for the previous “Appendix,” the notation here only pertains to this calculation and should not be confused with variables within the main text. It is convenient to translate (33) via \(B_2:=b_2-\xi \), to re-label \(x_2=X_2, y_2=Y_2\) and change to the timescale \(\tau =s/\epsilon ^2\) which yields

$$\begin{aligned} \begin{aligned} \dot{X}_2&= 3a_2(B_2+\xi )Y_2-X_2^2+B_2X_2+\delta ,\\ \dot{a}_2&= \epsilon ^2(\mu -\alpha a_2 -a_2(B_2+\xi )Y_2),\\ \dot{B}_2&= \epsilon ^2\epsilon _b(1-(B_2+\xi )X_2 -a_2(B_2+\xi )Y_2),\\ \dot{\epsilon }&= 0,\\ \dot{\delta }&= 0,\\ \dot{Y}_2&= \kappa (X_2^2-Y_2-a_2(B_2+\xi )Y_2).\\ \end{aligned} \end{aligned}$$
(82)

The system (82) has a line of equilibrium points

$$\begin{aligned} {\mathcal {E}}_2:=\{(X_2,a_2,B_2,\epsilon ,\delta ,Y_2)=(0,a_2,0,0,0,0)\}, \end{aligned}$$

which is degenerate since the linearization of (82) at \({\mathcal {E}}_2\) for fixed \(a_2=a_0\) is

$$\begin{aligned} A=\left. D_{(X_2,a_2,b_2,\epsilon ,\delta ,Y_2)} \left( \begin{array}{c} \dot{X}_2 \\ \dot{a}_2 \\ \dot{B}_2 \\ \dot{\epsilon } \\ \dot{\delta } \\ \dot{Y}_2\\ \end{array}\right) \right| _{{\mathcal {E}}_2}= \left( \begin{array}{cccccc} 0 &{} 0 &{} 0 &{} 0 &{} 1 &{} 3a_0\xi \\ 0 &{} 0 &{} 0 &{} \mu -\alpha a_0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} \epsilon _b &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0\\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} -\kappa (1+a_0\xi )\\ \end{array} \right) . \end{aligned}$$
(83)

This matrix has one negative eigenvalue \(-\kappa (1+a_0\xi )\) and a quintuple zero eigenvalue. Hence a center manifold reduction to a five-dimensional center flow is required to resolve the dynamics near \({\mathcal {E}}_2\). However, we use a preliminary transformation to get the system into standard form. Let \(Z=(X_2,Y_2)^T\) and set

$$\begin{aligned} A_{XY}:=\left( \begin{array}{cc} 0 &{} 3a_0\xi \\ 0 &{} -\kappa (1+a_0\xi ) \\ \end{array} \right) \quad \text {and} \quad M:=\left( \begin{array}{cc} 1 &{} -\frac{3 a_0 \xi }{\kappa (1 + a_0 \xi )} \\ 0 &{} 1 \\ \end{array} \right) . \end{aligned}$$

Then consider new coordinates via \(M\tilde{Z}=Z\) and observe that in the coordinates \((\tilde{X}_2,\tilde{Y}_2)^T\) we have

$$\begin{aligned} \tilde{Z}'=M^{-1}A_{XY}M\tilde{Z}+\text {h.o.t.}=\left( \begin{array}{cc} 0 &{} 0 \\ 0 &{} -\kappa (1+a_0\xi ) \\ \end{array} \right) \left( \begin{array}{c} \tilde{X}_2\\ \tilde{Y}_2 \\ \end{array} \right) +\text {h.o.t.}, \end{aligned}$$

where \(\text {h.o.t.}\) denotes higher-order terms. Let \(x:=(x_1,x_2,x_3,x_4,x_5)=(\tilde{X}_2,a_0-a_2,b_2,\epsilon ,\delta )\) so that \(y=\tilde{Y}_2\) is the transformation of (82) into new coordinates

$$\begin{aligned} \begin{aligned} x_1'&= \textstyle x_5 + 3 x_3 y (x_3 + \xi ) + x_3 \left( x_1 - \frac{3 a_0 y \xi }{\kappa + a_0 \kappa \xi } \right) - \left( x_1 - \frac{3 a_0 y \xi }{\kappa + a_0 \kappa \xi }\right) ^2\\&\quad \,\,- \textstyle 3 a_0 \xi \frac{y + (a_0-x_2) y (x_3 + \xi ) - \left( x_1 - \frac{3 a_0 y \xi }{\kappa + a_0 \kappa \xi }\right) ^2}{1 + a_0 \xi },\\ x_2'&= x_4 (\mu - (a_0-x_2) (\alpha + y (x_3 + \xi ))),\\ x_3'&= \textstyle x_4 \epsilon _b (1 - (a_0-x_2) y (x_3 + \xi ) - (x_3 + \xi ) \left( x_1 - \frac{3 a_0 y \xi }{\kappa + a_0 \kappa \xi }\right) ,\\ x_4'&=0,\\ x_5'&=0,\\ y'&= \textstyle \kappa \left( -y - (a_0-x_2) y (x_3 + \xi ) + \left( x_1 - \frac{3 a_0 y \xi }{\kappa + a_0 \kappa \xi }\right) ^2\right) , \end{aligned} \end{aligned}$$

which is a vector field we denote by \((Cx+F(x,y),Py+G(x,y))^T\) for \(F(x,y)\in {\mathbb {R}}^5, G(x,y)\in {\mathbb {R}}\). Observe that

$$\begin{aligned} C=\{A_{ij}\}_{i,j=1}^5,\quad \text {and}\quad P= -\kappa (1+a_0\xi ). \end{aligned}$$

The vector field is now in the correct form for center manifold theory, applied along the entire line of points parametrized by \(a_0\). The ansatz is

$$\begin{aligned} y=h(x)=\sum _{i+j=2,i\le j}c_{ij}x_ix_j. \end{aligned}$$

The usual invariance equation is given by

$$\begin{aligned} Dh(x)[Cx+F(x,h(x))]=Ph(x)+G(x,h(x)), \end{aligned}$$

where different powers \(x_ix_j\) have to have equal coefficients on both sides. This procedure yields

$$\begin{aligned} c_{11}=\frac{1}{1+a_0\xi },\quad c_{15}=-\frac{1}{\kappa (1+a_0\xi )^2}, \quad c_{55}=-\frac{1}{\kappa ^2(1+a_0\xi )^3}. \end{aligned}$$

All other coefficients \(c_{ij}\) have vanish. Hence, the center manifold is given to lowest order by

$$\begin{aligned} \tilde{Y}_2=\frac{\tilde{X}_2^2}{1+a_0\xi } -\frac{\tilde{X}_2\delta }{\kappa (1+a_0\xi )^2}-\frac{\delta ^2}{\kappa ^2(1+a_0\xi )^3}. \end{aligned}$$
(84)

Transforming back to original coordinates and keeping lowest order terms yields

$$\begin{aligned} Y_2=\frac{X_2^2}{1+a_0\xi }-\frac{\delta X_2}{\kappa (1+a_0\xi )^2} +\frac{\delta ^2}{\kappa ^2(1+a_0\xi )^3}+{\mathcal {O}}(Y_2^2,X_2^3,X_2Y_2,\delta Y_2,\delta ^3). \end{aligned}$$
(85)

Substituting the result into (82) gives, up to leading order, the center flow

$$\begin{aligned} \epsilon ^2\frac{dX_2}{ds}\!&= \! X_2^2\left( \frac{2a_0\xi -1}{1+a_0\xi }\right) \!+\!X_2\left( \frac{-\delta }{\kappa (1+a_0\xi )^2}+B_2\right) \!+\!\delta +\frac{\delta ^2}{\kappa ^2(1+a_0\xi )^3}+{\mathcal {O}}(3),\nonumber \\ \frac{da_2}{ds}&= \mu -\alpha a_2+{\mathcal {O}}(2),\\ \frac{dB_2}{ds}&= \epsilon _b+{\mathcal {O}}(2),\nonumber \end{aligned}$$
(86)

which is precisely the result we wanted to prove.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kuehn, C., Szmolyan, P. Multiscale Geometry of the Olsen Model and Non-classical Relaxation Oscillations. J Nonlinear Sci 25, 583–629 (2015). https://doi.org/10.1007/s00332-015-9235-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-015-9235-z

Keywords

Navigation