Skip to main content
Log in

Oscillations and Pattern Formation in a Slow–Fast Prey–Predator System

Bulletin of Mathematical Biology Aims and scope Submit manuscript

Cite this article

A Correction to this article was published on 28 October 2021

This article has been updated

Abstract

We consider the properties of a slow–fast prey–predator system in time and space. We first argue that the simplicity of the prey–predator system is apparent rather than real and there are still many of its hidden properties that have been poorly studied or overlooked altogether. We further focus on the case where, in the slow–fast system, the prey growth is affected by a weak Allee effect. We first consider this system in the non-spatial case and make its comprehensive study using a variety of mathematical techniques. In particular, we show that the interplay between the Allee effect and the existence of multiple timescales may lead to a regime shift where small-amplitude oscillations in the population abundances abruptly change to large-amplitude oscillations. We then consider the spatially explicit slow–fast prey–predator system and reveal the effect of different timescales on the pattern formation. We show that a decrease in the timescale ratio may lead to another regime shift where the spatiotemporal pattern becomes spatially correlated, leading to large-amplitude oscillations in spatially average population densities and potential species extinction.

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.

Institutional subscriptions

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

Similar content being viewed by others

Change history

Notes

  1. Data are taken from the Web of Science.

References

  • Allesina S, Bodini A (2004) Who dominates whom in the ecosystem? Energy flow bottlenecks and cascading extinctions. J Theor Biol 230:351–358

    Article  MathSciNet  MATH  Google Scholar 

  • Arditi R, Ginzburg LR (1989) Coupling in predator-prey dynamics: ratio-dependence. J Theor Biol 139:311–326

    Article  Google Scholar 

  • Baurmann M, Gross T, Feudel U (2007) Instabilities in spatially extended predator-prey systems: spatio-temporal patterns in the neighborhood of turing-Hopf bifurcations. J Theor Biol 245:220–229

    Article  MathSciNet  MATH  Google Scholar 

  • Benoit E, Callot JF, Diener F, Diener M (1981) Chasse au canard. Collectanea Mathematica 31–32:37–119

    MathSciNet  MATH  Google Scholar 

  • Brown JH (1994) Complex ecological systems. In: Cowan GA, Pines D, Melzer D (eds) Complexity: metaphors, models, and reality, Santa Fe Institute studies in the Science of Complexity, Proceedings, vol XVIII, Addison-Wesley, Reading, pp 419–449

  • Courchamp F, Berec L, Gascoigne J (2008) Allee effects in ecology and conservation. Oxford University Press, Oxford

    Book  Google Scholar 

  • Dennis B (1989) Allee effects: population growth, critical density, and the chance of extinction. Nat Resour Model 3:481–538

    Article  MathSciNet  MATH  Google Scholar 

  • Dumortier F (1978) Singularities of Vector Fields. IMPA, Rio de Janeiro, Brazil

  • Dumortier F (1993) Techniques in the theory of local bifurcations: blow-up, normal forms, nilpotent bifurcations, singular perturbations. In: Bifurcations and periodic orbits of vector fields, Springer, pp 19–73

  • Dumortier F, Roussarie R (1996) Canard cycles and center manifolds. Memoirs Am Math Soc 121:577

    Article  MathSciNet  MATH  Google Scholar 

  • Edwards AM, Brindley J (1999) Zooplankton mortality and the dynamical behaviour of plankton population models. Bull Math Biol 61:303–339

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Freedman HI (1980) Deterministic mathematical models in population ecology. Marcel Dekker, New York

    MATH  Google Scholar 

  • Gurney WSC, Veitch AR, Cruickshank I, McGeachin G (1998) Circles and spirals: population persistence in a spatially explicit predator–prey model. Ecology 79:2516–2530

    Google Scholar 

  • Gyllenberg M Personal communication

  • Hastings A, Powell T (1991) Chaos in a three-species food chain. Ecology 72:896–903

    Article  Google Scholar 

  • Hastings A, Harrison S, McCann K (1997) Unexpected spatial patterns in an insect outbreak match a predator diffusion model. Proc R Soc B 264:1837–1840

    Article  Google Scholar 

  • Hek G (2010) Geometric singular perturbation theory in biological practice. J Math Biol 60:347–386

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Sigmund K (1998) Evolutionary games and population dynamics. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Holling CS (1965) The functional response of predators to prey density and its role in mimicry and population regulation. Memoirs Entomol Soc Can 97(S45):5–60

    Article  Google Scholar 

  • Holt RD, Polis GA (1997) A theoretical framework for intraguild predation. Am Nat 149:745–764

    Article  Google Scholar 

  • Huang J, Ruan S, Song J (2014) Bifurcations in a predator–prey system of Leslie type with generalized Holling type III functional response. J Differ Equ 257:1721–1752

    Article  MathSciNet  MATH  Google Scholar 

  • Jankovic M, Petrovskii S (2014) Are time delays always destabilizing? Revisiting the role of time delays and the Allee effect. Theor Ecol 7:335–349

    Article  Google Scholar 

  • Jansen VAA (1995) Regulation of predator-prey systems through spatial interactions: a possible solution to the paradox of enrichment. Oikos 74:384–390

    Article  Google Scholar 

  • Jiang X, She Z, Ruan S (2021) Global dynamics of a predator–prey system with density-dependent mortality and ratio-dependent functional response. Discrete Continu Dyn Syst B 26:1967–1990

    Article  MathSciNet  MATH  Google Scholar 

  • Jordan F, Scheuring I, Vida G (2002) Species positions and extinction dynamics in simple food webs. J Theor Biol 215:441–448

    Article  MathSciNet  Google Scholar 

  • Kareiva PM (1990) Population dynamics in spatially complex environments: theory and data. Philos Trans R Soc B 330:175–190

    Article  Google Scholar 

  • Kooi BW, Poggiale JC (2018) Modelling, singular perturbation and bifurcation analyses of bitrophic food chains. Math Biosci 301:93–110

    Article  MathSciNet  MATH  Google Scholar 

  • Krupa M, Szmolyan P (2001a) Extending geometric singular perturbation theory to nonhyperbolic points- folds and canards in two dimension. SIAM J Math Anal 33(2):286–314

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

  • Kuehn C (2015) Multiple time scale dynamics. Springer, New York

    Book  MATH  Google Scholar 

  • Kuznetsov YA (2004) Elements of applied bifurcation theory. Springer, New York

    Book  MATH  Google Scholar 

  • Lewis MA, Kareiva P (1993) Allee dynamics and the spread of invading organisms. Theor Popul Biol 43:141–158

    Article  MATH  Google Scholar 

  • Lewis MA, Petrovskii S, Potts JR (2016) The mathematics behind biological invasions. Springer, New York

    Book  MATH  Google Scholar 

  • Ludwig D, Jones DD, Holling CS (1978) Qualitative analysis of insect outbreak systems: the spruce budworm and forest. J Anim Ecol 47:315–332

    Article  Google Scholar 

  • Malchow H, Petrovskii SV (2002) Dynamical stabilization of an unstable equilibrium in chemical and biological systems. Math Comput Model 36:307–319

    Article  MathSciNet  MATH  Google Scholar 

  • May RM (1972) Limit cycles in predator–prey communities. Science 177:900–902

    Article  Google Scholar 

  • McCauley E, Murdoch WW (1990) Predator–prey dynamics in environments rich and poor in nutrients. Nature 343:455–457

    Article  Google Scholar 

  • Medvinsky A, Petrovskii S, Tikhonova I, Malchow H, Li BL (2002) Spatiotemporal complexity of plankton and fish dynamics. SIAM Rev 44(3):311–370

    Article  MathSciNet  MATH  Google Scholar 

  • Mimura M, Murray JD (1978) On a diffusive prey–predator model which exhibits patchiness. J Theor Biol 75:249–262

    Article  MathSciNet  Google Scholar 

  • Morozov A, Petrovskii S, Li BL (2006) Spatiotemporal complexity of patchy invasion in a predator–prey system with the Allee effect. J Theor Biol 238:18–35

    Article  MathSciNet  MATH  Google Scholar 

  • Muratori S, Rinaldi S (1989) Remarks on competitive coexistence. SIAM J Appl Math 49(5):1462–1472

    Article  MathSciNet  MATH  Google Scholar 

  • Muratori S, Rinaldi S (1992) Low and high frequency oscillation in three-dimensional food chain systems. SIAM J Appl Math 52(6):1688–1706

    Article  MathSciNet  MATH  Google Scholar 

  • Murray JD (1968) Singular perturbations of a class of nonlinear hyperbolic and parabolic equations. J Math Phys 47:111–133

    Article  MathSciNet  MATH  Google Scholar 

  • Murray JD (1975) Non-existence of wave solutions for a class of reaction diffusion equations given by the Volterra interacting-population equations with diffusion. J Theor Biol 52(2):459–469

    Article  MathSciNet  Google Scholar 

  • Murray JD (1976) On travelling wave solutions in a model for the Belousov–Zhabotinskii reaction. J Theor Biol 56(2):329–353

    Article  MathSciNet  Google Scholar 

  • Murray JD (1981) A pre-pattern formation mechanism for animal coat marking. J Theor Biol 88:161–199

    Article  MathSciNet  Google Scholar 

  • Murray JD (1982) Parameter space for Turing instability in reaction–diffusion mechanisms: a comparison of models. J Theor Biol 98:143–163

    Article  MathSciNet  Google Scholar 

  • Murray JD (1988) How the leopard gets its spots. Sci Am 258:80–87

    Article  Google Scholar 

  • Murray JD (1989) Mathematical biology. Springer, New York

    Book  MATH  Google Scholar 

  • Pascual M (1993) Diffusion-induced chaos in a spatial predator–prey system. Proc R Soc B 251:1–7

    Article  Google Scholar 

  • Petrovskii SV, Malchow H (1999) A minimal model of pattern formation in a prey–predator system. Math Comput Model 29:49–63

    Article  MathSciNet  MATH  Google Scholar 

  • Petrovskii SV, Malchow H (2000) Critical phenomena in plankton communities: KISS model revisited. Nonlinear Anal Real World Appl 1(1):37–51

    Article  MathSciNet  MATH  Google Scholar 

  • Petrovskii SV, Malchow H (2001) Wave of chaos: new mechanism of pattern formation in spatio-temporal population dynamics. Theor Popul Biol 59(2):157–174

    Article  MATH  Google Scholar 

  • Petrovskii S, Kawasaki K, Takasu F, Shigesada N (2001) Diffusive waves, dynamical stabilization and spatio-temporal chaos in a community of three competitive species. Jpn J Ind Appl Math 18:459–481

    Article  MathSciNet  MATH  Google Scholar 

  • Petrovskii S, Vinogradov ME, Morozov A (2002) Formation of the patchiness in the plankton horizontal distribution due to biological invasion in a two-species model with account for the Allee effect. Oceanology 42:363–372

  • Poggiale JC, Aldebert C, Girardot B, Kooi BW (2020) Analysis of a predator–prey model with specific time scales: a geometrical approach proving the occurrence of canard solutions. J Math Biol 80:39–60

    Article  MathSciNet  MATH  Google Scholar 

  • Polis GA, Strong DR (1996) Food web complexity and community dynamics. Am Nat 147:813–846

    Article  Google Scholar 

  • Rinaldi S, Muratori S (1992) Slow-fast limit cycles in predator–prey models. Ecol Model 61:287–308

    Article  MATH  Google Scholar 

  • Rodrigues VW, Mistro DC, Rodrigues LAD (2020) Pattern formation and bistability in a generalist predator–prey model. Mathematics 8:20

    Article  Google Scholar 

  • Rosenzweig ML (1971) Paradox of enrichment: destabilization of exploitation ecosystem in ecological time. Science 171:385–387

    Article  Google Scholar 

  • Rosenzweig ML, MacArthur R (1963) Graphical representation and stability conditions of predator–prey interaction. Am Nat 97:209–223

    Article  Google Scholar 

  • Rovinsky A, Menzinger M (1992) Interaction of Turing and Hopf bifurcation in chemical systems. Phys Rev A 46(10):6315–6322

    Article  MathSciNet  Google Scholar 

  • Saha T, Pal PJ, Banerjee M (2021) Relaxation oscillation and canard explosion in a slow-fast predator–prey model with Beddington–DeAngelis functional response. Nonlinear Dyn 103:1195–1217

    Article  Google Scholar 

  • Scheffer M, Rinaldi S, Kuznetsov YA, Van Nes EH (1997) Seasonal dynamics of Daphnia and algae explained as a periodically forced predator–prey system. Oikos 80:519–532

    Article  Google Scholar 

  • Scheffer M, Rinaldi S, Kuznetsov YA (2000) Effects of fish on plankton dynamics: a theoretical analysis. Can J Fish Aquat Sci 57(6):1208–1219

    Article  Google Scholar 

  • Segel LA, Jackson JL (1972) Dissipative structure: an explanation and an ecological example. J Theor Biol 37:545–559

    Article  Google Scholar 

  • Sen M, Banerjee M, Morozov A (2011) Bifurcation analysis of a ratio-dependent prey–predator model with the Allee effect. Ecol Complex 11:12–27

    Article  Google Scholar 

  • Sen D, Petrovskii S, Ghorai S, Banerjee M (2020) Rich bifurcation structure of prey–predator model induced by the Allee effect in the growth of generalist predator. Int J Bifurc Chaos 30:2050084

    Article  MathSciNet  MATH  Google Scholar 

  • Sherratt JA (1998) Invading wave fronts and their oscillatory wakes are linked by a modulated travelling phase resetting wave. Physica D 117(1–4):145–166

    Article  MathSciNet  MATH  Google Scholar 

  • Sherratt JA, Lewis MA, Fowler A (1995) Ecological chaos in the wake of invasion. Proc Natl Acad Sci USA 92:2524–2528

    Article  MATH  Google Scholar 

  • Sherratt JA, Eagan BT, Lewis MA (1997) Oscillations and chaos behind predator–prey invasion: mathematical artifact or ecological reality? Philos Trans R Soc B 352:21–38

    Article  Google Scholar 

  • Siteur K, Eppinga MB, Doelman A, Siero E, Rietkerk M (2016) Ecosystems off track: rate-induced critical transitions in ecological models. Oikos 125:1689–1699

    Article  Google Scholar 

  • Song YL, Zhang TH, Peng YH (2016) Turing-Hopf bifurcation in the reaction–diffusion equations and its applications. Commun Nonlinear Sci Numer Simul 33:229–258

    Article  MathSciNet  MATH  Google Scholar 

  • Stenseth NC, Falck W, Bjornstad ON, Krebs CJ (1997) Population regulation in snowshoe hare and Canadian lynx: asymmetric food web configurations between hare and lynx. Proc Natl Acad Sci USA 94(10):5147–5152

    Article  Google Scholar 

  • Stephens PA, Sutherland WJ (1999) Consequences of the Allee effect for behaviour, ecology and conservation. Trends Ecol Evol 14(10):401–405

    Article  Google Scholar 

  • Tu LW (2008) An introduction to manifolds. Springer, New York

    MATH  Google Scholar 

  • Turchin P (2003) Complex population dynamics: a theoretical/empirical synthesis. Princeton University Press, Princeton

    MATH  Google Scholar 

  • Turing AM (1952) On the chemical basis of morphogenesis. Philos Trans R Soc B 237:37–72

    MathSciNet  MATH  Google Scholar 

  • Van der Pol B (1926) On “relaxation-oscillations’’. Lond Edinb Dublin Philos Mag J Sci Ser 7(2):978–992

    Google Scholar 

  • Volterra V (1926) Fluctuations in the abundance of a species considered mathematically. Nature 118:558–560

    Article  MATH  Google Scholar 

  • Wang C, Zhang X (2019a) Canards, heteroclinic and homoclinic orbits for a slow-fast predator-prey model of generalized Holling type IIIs. J Differ Equ 267:3397–3441

  • Wang C, Zhang X (2019b) Relaxation oscillations in a slow-fast modified Leslie–Gower model. Appl Math Lett 87:147–153

  • Zou R, Guo S (2020) Dynamics of a diffusive Leslie-Gower predator-prey model in spatially heterogeneous environment. Discrete Continu Dyn Syst B 25:4189–4210

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This paper has been supported by the RUDN University Strategic Academic Leadership Program (to S.P.).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Malay Banerjee.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

The original article was revised: The notation of \(\lambda _H\) was shown incorrectly in a few places in the published version. It has been corrected.

Appendices

Appendix A

Here, we will follow geometric singular perturbation technique as given by Fenichel (1979) to find the analytical expression of locally perturbed invariant manifold \(C^1_{\varepsilon }\). Since \(v=q(u,\varepsilon )\), from the invariance condition we have

$$\begin{aligned} \dfrac{\hbox {d}v}{\hbox {d}t} = \dfrac{\hbox {d}q(u,\varepsilon )}{\hbox {d}u}\dfrac{\hbox {d}u}{\hbox {d}t}. \end{aligned}$$

Using the explicit expression for \(\dfrac{\hbox {d}u}{\hbox {d}t}\) and \(\dfrac{\hbox {d}v}{\hbox {d}t}\) from (2), we get

$$\begin{aligned} \begin{aligned} \varepsilon q(u,\varepsilon )(u(1-\alpha \delta )-\delta ) = u \dfrac{\hbox {d}q(u,\varepsilon )}{\hbox {d}u}(\gamma (1-u)(u+\beta )(1+\alpha u)-q(u,\varepsilon )). \end{aligned} \end{aligned}$$
(35)

Substituting the asymptotic expansion of \(q(u,\varepsilon )\) from (9 and assuming \(u\ne 0\), \(\dot{q_0}(u)\ne 0,\) we equate \(\varepsilon \) free terms from both sides to obtain

$$\begin{aligned} \begin{aligned} q_0(u)&= \gamma (1-u)(u+\beta )(1+\alpha u),\\ \end{aligned} \end{aligned}$$
(36)

which is exactly the critical manifold. Now equating the coefficients of \(\varepsilon \) from both sides of (35) we get

$$\begin{aligned} \begin{aligned} q_1(u) = \dfrac{q_0(u)(u(1-\alpha \delta )-\delta )}{-\,u{\dot{q}}_0(u)}. \end{aligned} \end{aligned}$$
(37)

Similarly, we obtain \(q_2(u)\) by equating the coefficients of \(\varepsilon ^2,\)

$$\begin{aligned} \begin{aligned} q_2(u) = \dfrac{q_1(u)(u(1-\alpha \delta )-\delta )+u q_1\dot{q_1}(u)}{-\,u\dot{q_0}(u)}. \end{aligned} \end{aligned}$$
(38)

Proceeding as above, we find \(q_r(u),\ r=3,4,\ldots \) by equating the coefficients of \(\varepsilon ^r\) from (35). Therefore, the second-order approximation of the perturbed invariant manifold is given by

$$\begin{aligned} q(u,\varepsilon )=q_0(u)+\varepsilon q_1(u)+\varepsilon ^2 q_2(u), \end{aligned}$$

where \(q_0,q_1,q_2\) are given in Eqs. (36)–(38).

Appendix B

We apply the blow-up transformation in the slow–fast normal form (12) where

$$\begin{aligned}&h_1(U,V)=u_*+U, \ \ h_3(U,V)=0,\ \ h_5(U,V)=(v_*+V)(1+\alpha u_*)+Uv_*\alpha , \\&h_2(U,V)=-\,\gamma (-1+6u_*^2\alpha +3u_*(1+\alpha (\beta -1))+\beta -\alpha \beta )-U\gamma (1+\alpha (4u_*+\beta -1)), \\&h_4(U,V)=(v_*+V)(1-\alpha \delta _*),\ \ h_6(U,V)=u_*-(1+u_*\alpha )\delta _*, \end{aligned}$$

\({\bar{\varepsilon }}=1\) the blow-up transformation as defined in (13) reduces to

$$\begin{aligned} {\bar{r}} = \sqrt{\varepsilon },\ U=\sqrt{\varepsilon }{\bar{U}},\ V=\varepsilon {\bar{V}},\ \lambda =\sqrt{\varepsilon }{\bar{\lambda }}. \end{aligned}$$
(39)

Using the transformation (39), we can write the system (15) by removing the overbars as

$$\begin{aligned} \begin{aligned} U_t&= -\,b_1V+b_2U^2+\sqrt{\varepsilon }{\mathcal {G}}_1(U,V) + O(\sqrt{\varepsilon }(\lambda +\sqrt{\varepsilon })),\\ V_t&= b_3U-b_4\lambda +\sqrt{\varepsilon }{\mathcal {G}}_2(U,V) + O(\sqrt{\varepsilon }(\lambda +\sqrt{\varepsilon })),\\ \end{aligned} \end{aligned}$$
(40)

where

$$\begin{aligned} \begin{aligned} \&b_1 =u_*,\ \ b_2 =-\,\gamma (-1+6u_*^2\alpha +3u_*(1+\alpha (\beta -1))+\beta -\alpha \beta ),\\ \&b_3 = v_*(1-\alpha \delta _*),\ \ b_4 = v_*(1+\alpha u_*), \end{aligned} \end{aligned}$$
(41)

and

$$\begin{aligned} \begin{aligned} {\mathcal {G}}_1(U,V) = a_1U-a_2UV+a_3U^3,\ \ {\mathcal {G}}_2(u,V) = a_4U^2+a_5V. \end{aligned} \end{aligned}$$
(42)

Let the equilibrium point of the system (40) is \((U_e,V_e)\), \(U_e=\dfrac{b_4\lambda }{b_3}+O(2)\) and \(V_e=O(2)\) where \(O(2):=O(\lambda ^2,\lambda \sqrt{\varepsilon },\lambda )\). Linearizing the system about this equilibrium point we have the Jacobian matrix as

$$\begin{aligned} {\mathcal {J}}:= \begin{pmatrix} 2U_eb_2+a_1 \sqrt{\varepsilon }+O(2)&{}-b_1+O(2)\\ b_3+O(2)&{}a_5\sqrt{\varepsilon }+O(2) \end{pmatrix}. \end{aligned}$$
(43)

At the Hopf bifurcation, we have Trace \({\mathcal {J}}=0\) which implies

$$\begin{aligned} \dfrac{2b_2b_4\lambda }{b_3}+\sqrt{\varepsilon }(a_1+a_5)+O(2)=0. \end{aligned}$$
(44)

and applying the blow-down map \({\lambda _\mathrm{H}}=\lambda \sqrt{\varepsilon }\) we get the singular Hopf bifurcation curve \({\lambda _\mathrm{H}}(\sqrt{\varepsilon })\) for the slow–fast normal form (12) as

$$\begin{aligned} {\lambda _\mathrm{H}}(\sqrt{\varepsilon })=-\,\dfrac{b_3(a_1+a_5)}{2b_2b_4}\varepsilon +O(\varepsilon ^{3/2}). \end{aligned}$$
(45)

Appendix C

Here, we prove the existence of maximal canard curve and will give an analytical expression for the same. For that we will first prove the following proposition. In chart \(K_2\) of the blow-up space, we consider the desingularized system (40) as

$$\begin{aligned} \begin{aligned} U_t&= -\,b_1V+b_2U^2+r{\mathcal {G}}_1(U,V) + O(\lambda r,r),\\ V_t&= b_3U-b_4\lambda +r{\mathcal {G}}_2(U,V) + O(\lambda r,r),\\ r_t&=0,\\ \lambda _t&=0, \end{aligned} \end{aligned}$$
(46)

where \(b_1,\ b_2,\ b_3,\ b_4, {\mathcal {G}}_1\) and \({\mathcal {G}}_2\) are computed above (41), (42). The dynamics of the system on the sphere is obtained by putting \(r=0\) in (46) for different values of \(\lambda \) in the vicinity of 0. Thus, by taking \(r=0,\ \lambda =0\), the above system is integrable and we have

$$\begin{aligned} \begin{aligned} U_t&= -\,b_1V+b_2U^2,\\ V_t&= b_3U. \end{aligned} \end{aligned}$$
(47)

This is a Riccati equation and the solution of this equation helps in proving our main theorem.

Proposition 1

The solution of the system (47) is given by \(H(U,V) = c,\) where

$$\begin{aligned} H(U,V) = e^{-\dfrac{2b_2}{b_3}V}\left( \dfrac{b_3}{2}U^2-\dfrac{b_1b_3^2}{4b_2^2}-\dfrac{b_1b_3}{2b_2}V\right) \end{aligned}$$

and

$$\begin{aligned}&\dfrac{\hbox {d}U}{\hbox {d}t} = -\,e^{\dfrac{2b_2}{b_3}V}\dfrac{\partial H}{\partial V}, \nonumber \\&\dfrac{\hbox {d}V}{\hbox {d}t} = e^{\dfrac{2b_2}{b_3}V}\dfrac{\partial H}{\partial U}. \end{aligned}$$
(48)

Proof

We can write the above Riccati system (47) as

$$\begin{aligned} \dfrac{\hbox {d}V}{\hbox {d}U}=\dfrac{b_3U}{-b_1V+b_2U^2} \end{aligned}$$
(49)

where the integrating factor is \(e^{-\dfrac{2b_2}{b_3}V}\). Multiplying both sides with the I.F and integrating we get

$$\begin{aligned} e^{-\dfrac{2b_2}{b_3}V} \left( U^2-\dfrac{b_1}{b_2}V-\dfrac{b_1b_3}{2b_2^2}\right) =c_0. \end{aligned}$$

Multiplying with \(\dfrac{b_3}{2}\), we obtain the solution of the system( 47) as

$$\begin{aligned} e^{-\dfrac{2b_2}{b_3}V}\left( \dfrac{b_3}{2}U^2-\dfrac{b_1b_3^2}{4b_2^2}-\dfrac{b_1b_3}{2b_2}V\right) = c, \end{aligned}$$

where \(c=c_0\dfrac{b_3}{2}\) is a constant. The solution determined by \(c=0\) is a parabola of the form

$$\begin{aligned} U^2=\dfrac{b_1b_3}{2b_2^2}+\dfrac{b_1}{b_2}V. \end{aligned}$$

\(\square \)

Proof of theorem 4.2:

We write the solution of the system (47) in the parametric form

$$\begin{aligned} \eta (t) = (U(t),V(t))= \left( t,\dfrac{b_2}{b_1}t^2-\dfrac{b_3}{2b_2}\right) ,\ t\in \mathbb {R} \end{aligned}$$
(50)

For \(\varepsilon =0\), the attracting and repelling submanifolds of the critical manifold \({\mathcal {M}}^1_0\) intersect along the equator of the blow-up space \(S^3\). From Fenichel’s theory, for \(\varepsilon >0\) there exist invariant perturbed attracting \(({\mathcal {M}}_{\varepsilon }^{1,a})\) and repelling submanifold \(({\mathcal {M}}_{\varepsilon }^{1,r})\). Along the curve (50), the attracting \(({\mathcal {M}}_{\varepsilon }^{1,a})\) and repelling \(({\mathcal {M}}_{\varepsilon }^{1,r})\) invariant submanifolds in the blow-up space intersect and the solution trajectory lying in that intersection is called maximal canard. We use Melnikov function to calculate the distance between these invariant manifolds (Krupa and Szmolyan 2001a; Kuehn 2015), which is given by

$$\begin{aligned} D_{r,\lambda } = d_r r + d_{\lambda } \lambda + O(r^2), \end{aligned}$$
(51)

where

$$\begin{aligned} \begin{aligned} d_r = \int _{-\infty }^{\infty }\nabla H(\eta (t))^T{\mathcal {G}}(\eta (t))\hbox {d}t,\\ d_{\lambda } = \int _{-\infty }^{\infty }\nabla H(\eta (t))^T\begin{pmatrix} 0\\ -b_4 \end{pmatrix}\hbox {d}t, \end{aligned} \end{aligned}$$
(52)

where \({\mathcal {G}},\ H\) and \(b_4\) are defined in (42), (48) and (41), respectively. The distance between the submanifolds \({\mathcal {M}}_{\varepsilon }^{1,a}\) and \({\mathcal {M}}_{\varepsilon }^{1,r}\) is given by Eq. (51). And since the maximal canard lie in the intersection of these manifolds, so we must have \(D_{r,\lambda }=0\). For that, we now calculate the Melnikov-type integrals \(d_r\) and \(d_{\lambda }\) (50 ) and (52). Therefore,

$$\begin{aligned} \begin{aligned} d_r&= \int _{-\infty }^{\infty }\left[ (a_1U-a_2UV+a_3U^3 )\dfrac{\partial H(\eta (t))}{\partial U}+(a_4U^2+a_5V)\dfrac{\partial H(\eta (t))}{\partial V}\right] \hbox {d}t\\&= \int _{-\infty }^{\infty }e^{-\dfrac{2b_2}{b_3}V}\left[ (a_1U-a_2UV+a_3U^3 )b_3U+(a_4U^2+a_5V)(b_1V-b_2U^2)\right] \hbox {d}t\\&= e\int _{-\infty }^{\infty }e^{-A_4t^2}\left( A_1t^4+A_2t^2+A_3\right) \hbox {d}t \end{aligned} \end{aligned}$$
(53)

where

$$\begin{aligned} A_1 = a_3b_3-\dfrac{a_2b_2b_3}{b_1},\ A_2 = a_1b_3+ \dfrac{a_2b_3^2}{2b_2}-\dfrac{a_4b_1b_3}{2b_2}-\dfrac{a_5b_3}{2},\ A_3 = \dfrac{a_5b_1b_3^2}{4b_2^2},\ A_4 = \dfrac{2b_2^2}{b_1b_3}. \end{aligned}$$

Now substituting \(z=t^2\) and by repeated integration by parts, we obtain

$$\begin{aligned} \begin{aligned} d_r = e\left( \dfrac{3A_1}{4A_4^2}+\dfrac{A_2}{2A_4}+A_3\right) \int _{-\infty }^{\infty }e^{-A_4t^2}\hbox {d}t, \end{aligned} \end{aligned}$$
(54)

and

$$\begin{aligned} \begin{aligned} d_{\lambda }&= -\,\int _{-\infty }^{\infty }b_4\dfrac{\partial H}{\partial V}\hbox {d}t\\&= b_4\int _{-\infty }^{\infty }e^{-\dfrac{2b_2}{b_3}V}(-b_1V+b_2U^2)\hbox {d}t\\&= e A_5\int _{-\infty }^{\infty }e^{-A_4t^2}\hbox {d}t, \end{aligned} \end{aligned}$$
(55)

where \(A_5 = \dfrac{b_1b_3b_4}{2b_2}\). Since \(d_{\lambda }\ne 0\) therefore using implicit function theorem we can explicitly solve for \(\lambda \) from (51)

$$\begin{aligned} \begin{aligned} \lambda (r)&=-\,\dfrac{\hbox {d}_r}{\hbox {d}_\lambda }r + O(r^2) = -\dfrac{1}{A_5}\left( \dfrac{3A_1}{4A_4^2}+\dfrac{A_2}{2A_4}+A_3\right) r + O(r^2). \end{aligned} \end{aligned}$$
(56)

Now using blow down map \(\lambda _c=\lambda \sqrt{\varepsilon }\), we obtain the maximal canard curve for the slow–fast normal form (12).

$$\begin{aligned} \begin{aligned} \lambda _c(\sqrt{\varepsilon }) = -\,\dfrac{1}{A_5}\left( \dfrac{3A_1}{4A_4^2}+\dfrac{A_2}{2A_4}+A_3\right) \varepsilon + O(\varepsilon ^{3/2}). \end{aligned} \end{aligned}$$
(57)

\(\square \)

Appendix D

Here, we prove the existence of a unique attracting limit cycle called relaxation oscillation. To study the dynamics of the system (21), we define two sections of the flow as

$$\begin{aligned} \begin{aligned}&\varDelta ^\mathrm{in}=\{(u_+,v):u_+<<u_\mathrm{max}, v\in (v_1-\rho ,v_1+\rho )\},\\&\varDelta ^\mathrm{out}=\{(u_+,v):u_+<<u_\mathrm{max}, v\in (v_0-\rho ^2,v_0+\rho ^2)\}, \end{aligned} \end{aligned}$$

where \(u_\mathrm{max},\ v_1,\ v_0\) are defined in Sect. 4.3 and \(\rho \) is sufficiently small positive number.

Let us define a return map \(\varPi :\varDelta ^\mathrm{in}\rightarrow \varDelta ^\mathrm{in}\) which is a composition of two maps

$$\begin{aligned} \varPhi :\varDelta ^\mathrm{in}\rightarrow \varDelta ^\mathrm{out},\ \ \varPsi :\varDelta ^\mathrm{out}\rightarrow \varDelta ^\mathrm{in}, \end{aligned}$$

such that \(\varPi = \varPsi \circ \varPhi \). Let us fix \(\varepsilon >0\) and we take a point \((u_+,v_+)\) on the section \(\varDelta ^\mathrm{in}\). Now we consider a trajectory of the system (21) starting from the initial point \((u_+,v_+)\). From the analysis of the entry–exit function, we can say that this trajectory will be attracted to \(V_+\) and will leave \(V_-\) at point \((0,p(v_+)),\) where p is the entry–exit function. The trajectory then jumps into the section \(\varDelta ^\mathrm{out}\) at the point \((u_+,p(v_+)).\) Thus, the map \(\varPhi \) is defined with the help of entry–exit function as \(\varPhi (u_+,v_+)=(u_+,p(v_+)).\)

Now to study the map \(\varPsi \) we consider two trajectories \(\gamma _{\varepsilon }^1, \gamma _{\varepsilon }^2\) starting from the section \(\varDelta ^\mathrm{out}\). These trajectories get attracted toward \(C_{\varepsilon }^{1,a}\) where the slow flow is given by \(\dfrac{\hbox {d}u}{\hbox {d}\tau }=\dfrac{g(u,q(u,\varepsilon ))}{{\dot{q}}(u,\varepsilon )}.\)

They follow the slow perturbed manifold until the vicinity of the fold point where they contract exponentially toward each other (Wang and Zhang 2019b) and jump into \(\varDelta ^{in}.\) From Theorem 2.1 of Krupa and Szmolyan (2001a), we have that the map \(\varPi \) is a contraction. Using contraction mapping theorem, we conclude that \(\varPi \) has a unique fixed point which gives rise to a unique relaxation oscillation cycle \(\gamma _{\varepsilon }\). Further from Fenichel’s theory, we infer that \(\gamma _{\varepsilon }\) converges to \(\gamma _0\) as \(\varepsilon \rightarrow 0.\)

For the parameter values \(\alpha =0.5,\ \beta =0.2,\ \delta =0.3,\) the unique attracting cycle \(\gamma _{\varepsilon }\) for \(\varepsilon =0.1\), is shown in Fig. 15 which converges to \(\gamma _0\) as \(\varepsilon \rightarrow 0\).

Fig. 15
figure 15

Singular trajectory \(\gamma _0\) (blue) and unique attracting limit cycle \(\gamma _{\varepsilon }\) for \(\varepsilon =0.1\) (green) for \(\alpha =0.5,\ \beta =0.2,\ \delta =0.3\) (Color figure online)

Rights and permissions

Reprints and Permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chowdhury, P.R., Petrovskii, S. & Banerjee, M. Oscillations and Pattern Formation in a Slow–Fast Prey–Predator System. Bull Math Biol 83, 110 (2021). https://doi.org/10.1007/s11538-021-00941-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-021-00941-0

Keywords

Navigation