286
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Codimension-one bifurcation analysis and chaos of a discrete prey–predator system

ORCID Icon &
Article: 2317505 | Received 12 May 2023, Accepted 07 Feb 2024, Published online: 28 Feb 2024

Abstract

In this paper, we explore local dynamics at equilibrium points, the existence of bifurcation sets and codimension-one bifurcation analysis of a discrete prey–predator system with Holling type-II functional response. Further, OGY and Hybrid control strategies are utilized to control chaos in the under study discrete model due to the occurrence of Neimark–Sacker and flip bifurcations. Finally, numerical simulations are given to verify theoretical results.

MSC classifications:

1. Introduction

1.1. Motivation and mathematical formulation

The Lotka–Volterra model, often known as the prey–predator model, is a mathematical representation of the dynamics of two species, one being the predator and the other being the prey in a given population. The model explains the long-term interconnections among the populace of the predator and prey species. It considers variables like the rate of birth and death of the predator and prey, as well as the rate at which the predator hunts and kills the prey. The prey–predator model has been broadly used to explore the dynamics of population interactions in various fields such as designing conservation procedure to safeguard endangered species and understanding the interconnections between several species in an ecosystem. Although these dynamical concerns resulting from the mathematical simulation of prey–predator systems may initially appear to be straightforward, a deep study of these models frequently give rise to incredibly challenging problems. The main goal of modelling a population ecosystem is to ensure that the mathematical model in question can represent the noteworthy system behaviours for the system being observed. Dynamic modelling of ecological systems is often an evolving technique. Prey–predator relationships modelled mathematically using the famous Lotka–Volterra type prey–predator model with Holling type-II functional response [Citation1]: (1) {x˙=ax(1xK)cxym+x,y˙=y(fxm+xd),(1) where x is prey's density and y is predator's density, while parameters c, a, d, f, K, and m show the capturing rate, intrinsic growth rate of prey, predator death rate, maximal predator growth rate, carrying capacity and half saturation constant, respectively. The so-called ratio-dependent theory, which holds that the rate of per capita predator growth should rely on the ratio of prey to predator abundance, should be the foundation of a more suitable general theory of predator–prey relationships. This is especially true when the predator has to look for food, and it is supported by numerous field, lab studies and observations [Citation2–5]. The basic form of ratio-dependent prey–predator system is: (2) {x˙=xf(x)yp(xy),y˙=(cq(xy)d)y,(2) where c is conversion rate, predator functional response is represented by p(x), q(x) is supplanted by p(x). Moreover, Kuang and Beretta [Citation6] have extended the prey–predator model (Equation2) by the effect of Holling type-II functional response as follows: (3) {x˙=ax(1xK)cxymy+x,y˙=y(d+fxmy+x).(3) It is anticipated that model (Equation3) becomes the form [Citation6, Citation7]: (4) {x˙=x(1x)sxyx+y,y˙=δy(r+xx+y),(4) by (5) {x=xK,t=at,y=myK,(5) where (6) {s=cma,δ=fa,r=df.(6)

1.2. Literature survey

Many investigators investigated the dynamics of prey–predator models. For instance, Singh and Malik [Citation8] have investigated the dynamics at equilibrium points and exploration the bifurcation of the following model: (7) {xt+1=xt+xt(1xtpytγ+xtαδ+xt),yt+1=yt+βyt(1qytγ+xt),(7) where yt and xt denote the predator and prey populace densities. Ma et al. [Citation9] have studied the local dynamics and bifurcation of following model: (8) {xt+1=xt+xt1+kyt(raxt)(1m)xtytb+(1m)xt,yt+1=yt(1+μcytb+(1m)xt),(8) where r,a,m,b,c,r,μ,k are positive constants. Santra et al. [Citation10] have explored behaviour of the following model involving prey refuge: (9) {xt+1=axt(1xt)c(xtbyt)yt,yt+1=d(xtbyt)yt,(9) where a, c, b, r, d are positive parameters. Chen et al. [Citation11] have studied local stability and bifurcation at fixed points of the following model: (10) {xt+1=xt+xt(1xt)xtu+xtbxtyt,yt+1=yt+cxtyt+mytyt2,(10) where u, b, c, m are positive constants. In contrast to discrete models, many investigators investigated the dynamics of continuous-time systems designated by differential equations. For instance, Tunç and Tunç [Citation12] have explored dynamics of the following system of second-order differential equations with delay: (11) x¨+F(x,x˙)x˙+H(x(tτ(t)))=P(t,x,x˙).(11) Tunç [Citation13] has investigated the dynamics of the following vector lienard equation: (12) x¨+F(x,x˙)x˙+H(x(tτ))=P(t).(12) Tunç [Citation14] has investigated the boundedness of solutions of a class of non-autonomous differential equations.

1.3. The novelty of the proposed work

It is important here to note that the earlier work is about the global qualitative study on system (Equation3) where the discussion of authors focuses on the qualitative study of the ratio-dependent predator–prey systems. Their work builds upon previous work but aims to address the open questions regarding the global qualitative behaviour of the model. By transforming the Michaelis-Menten-type ratio-dependent model into a Gause-type predator–prey system, the authors obtain a complete classification of the asymptotic behaviour of the solutions. They resolve open questions regarding the global stability of equilibria and the uniqueness of limit cycles. Their work provides insights into how the outcomes of the model depend on initial conditions. Furthermore, they present the biological implications of the findings, showcasing the practical relevance of the study. Overall, Kuang and Beretta [Citation6] discussed the dynamics of continuous-time model (Equation3), and demonstrate that if the system's positive steady state (Equation3) is locally asymptotically stable, then the system cannot have non-trivial positive periodic solutions. It also includes some results regarding the global stability of the positive steady state. Moreover, they non-dimensionalize the system (Equation3) into (Equation4), then, Hsu et al. [Citation7] investigated the full classification for the asymptotic behaviour of (Equation4) by transforming it into a Gause-type predator–prey system.

In contrast to the continuous-time model, our aim in this paper is to explore the dynamical characteristics of the following prey–predator system: (13) {xt+1=xt(1+h)1+hxthsxtyt(xt+yt)(1+hxt),yt+1=yt+yt(r+xtxt+yt),(13) which is discrete version of (Equation4) by non-standard finite difference scheme, where h is a step size. Furthermore, s denotes the ratio of capturing rate to the half saturation constant times intrinsic growth rate of prey, δ is the ratio of maximal predator growth rate to the intrinsic growth rate of prey, and finally r is the ratio of predator death rate to the maximal predator growth rate. The reason for studying the dynamics of discrete-time system (Equation13), instead of of continuous-time system (Equation4), is that the discrete model simplifies the complexity of the system, making it more accessible for analysis and interpretation. Continuous models often involve differential equations, which can be challenging to solve analytically or numerically. By converting the model into a set of difference equations, the computational complexity is reduced, making it easier to simulate and analyse the dynamics of the system. Additionally, the discretization of the model allows for direct comparison with experimental data. Experimental data is often collected at discrete-time points, and by converting the model into a discrete version, the model's predictions can be directly compared with the observed data. This facilitates experimental validation and enhances the credibility of the model. Furthermore, stability analysis becomes more feasible with discrete models. By studying the behaviour of the system over discrete-time steps, it becomes possible to analyse the stability of the fixed points and determine if the system will converge to an equilibrium or exhibit oscillatory behaviour. This stability analysis can provide valuable insights for system design and control. Lastly, parameter estimation is often easier with discrete models. Estimating parameters in continuous models can be challenging due to the complexity of the equations. Discrete models, with their simpler forms, allow for more straightforward parameter estimation techniques, enhancing the accuracy and reliability of the model's predictions. In summary, converting a continuous-time model into a discrete version offers advantages in terms of computational efficiency, simplicity, compatibility with experimental data, stability analysis, and parameter estimation. These advantages make the discrete model a valuable tool in the literature of ecology and biology for studying and analysing the local behaviour of fixed points. In the context of ecology and biology, calculating the local behaviour for fixed points in discrete models offers several advantages. Firstly, it allows for the identification of bifurcation sets, which are critical points where the qualitative behaviour of the system changes. By analysing the local behaviour of fixed points, researchers can identify bifurcations such as the emergence of multiple stable states or the occurrence of oscillatory behaviour. Understanding these bifurcation sets provides insights into the underlying mechanisms driving system dynamics. Secondly, discrete-time models enable bifurcation analysis, which involves studying how the system's behaviour changes as model parameters are varied. This analysis can reveal the existence of different types of bifurcations, such as saddle-node, transcritical, Neimark–Sacker, flip, or pitchfork bifurcations. Bifurcation analysis helps in understanding the stability and dynamics of the system under different conditions. Additionally, discrete models are useful for studying chaos in ecological and biological systems. Techniques like the OGY (Ott-Grebogi-Yorke) method can be employed to analyse the local behaviour of fixed points and identify chaotic regions in the parameter space. This study of chaos can provide insights into the underlying mechanisms driving chaotic behaviour. Furthermore, discrete-time models facilitate the development and analysis of hybrid control strategies. Hybrid control combines continuous and discrete dynamics to design control strategies that effectively manage complex systems. By calculating the local behaviour for fixed points, researchers can determine control strategies that stabilize the system, suppress undesirable behaviour, or guide the system towards desired states. Lastly, discrete-time models offer opportunities for numerical validation of theoretical results. Researchers can simulate the discrete model and compare the results with theoretical predictions, validating the accuracy and reliability of their theoretical findings. This numerical validation helps build confidence in the model and its ability to capture the essential dynamics of the system. In summary, calculating the local behaviour for fixed points in discrete models provides advantages in identifying bifurcation sets, conducting bifurcation analysis, studying chaos, developing hybrid control strategies, and numerically validating theoretical results. These advantages enhance our understanding of ecological and biological systems and enable more effective decision-making and control strategies. Due to aforementioned fact, in this paper we explore dynamics of discrete system model (Equation13) where our key investigations include:

  • Local behaviour for fixed points.

  • Identification of bifurcation sets for fixed points.

  • Bifurcation analysis of discrete model (Equation13).

  • Study of chaos by OGY and Hybrid control strategies.

  • Numerical validation of theoretical results.

1.4. The advantages of the proposed work

In the setting of a discrete prey–predator system, codimension-one bifurcation analysis has numerous advantages for understanding the dynamics, stability, and potential for chaos in ecological systems:

  • Codimension-one bifurcation analysis aids in identifying key parameter value at which qualitative changes in the system dynamics occur. It enables researchers to identify the precise parameter value that cause bifurcations or the formation of chaotic attractors.

  • This type of analysis gives information about the stability of the system's equilibrium points and periodic orbits. It helps identify whether prey and predator populations tend to stabilize or oscillate under varied parametric value.

  • Researchers can forecast system transitions by examining codimension-one bifurcations. They can, for example, forecast when a stable equilibrium will become unstable, resulting in the formation of chaotic behaviour.

  • Codimension-one bifurcation analysis can aid in detecting the emergence of chaos in a discrete prey–predator system.

  • Codimension-one bifurcation analysis can be used as a teaching tool for students and researchers to learn about complex ecological dynamics.

  • The calculation of local behaviour for fixed points in discrete-time models offers advantages in identifying bifurcation sets, conducting bifurcation analysis, studying chaos, developing hybrid control strategies, and numerically validating theoretical results. These advantages enhance our understanding of ecological systems, aid in predicting and managing ecological dynamics, and inform conservation and ecosystem management practices.

1.5. Paper layout

The organization of rest of the paper is as follows: stability analysis and bifurcation sets are studied in Section 2. In Section 3, we studied flip bifurcation for boundary fixed point, whereas Neimark–Sacker and flip bifurcations at interior fixed points are briefly investigated in Sections 4 and 5, respectively. The chaos control by OGY and Hybrid control strategies is presented in Section 6. Finally, numerical simulations and conclusion are given in Sections 7 and 8, respectively.

2. Bifurcation sets and stability analysis

In the present section, we explore the stability analysis at equilibrium points and bifurcation sets for the discrete prey–predator model (Equation13). If φ=(x,y) is an equilibrium point of (Equation13), then one has (14) {x=x(1+h)1+hxhsxy(x+y)(1+hx),y=y+hδy(r+xx+y).(14) Since φ1=(1,0) satisfied system (Equation14) obviously, and therefore for all h, s, δ and r discrete model (Equation13) has semitrivial equilibrium point φ1=(1,0). In order to get the model's interior equilibrium point, one need to solve the following system (15) {1+h1+hxhsy(x+y)(1+hx)=1,δ(r+xx+y)=0.(15) Equation (Equation15) can also be rewritten as (16) y=x2x1xs,(16) and (17) y=(1r)xr.(17) Using (Equation17) into (Equation16), one gets (18) x=1s+sr.(18) Further, using (Equation18) into (Equation17), one gets (19) y=(1r)(1s+sr)r.(19) So, from Equations (Equation18) and (Equation19), one can obtained that if s1s<r<1 then in R+2 model's interior equilibrium point is φ2=(1s+sr,(1r)(1s+sr)r). Now the variation matrix V|φ of the linearized system of (Equation13) under the map (f1,f2)(xt+1,yt+1) is (20) V|φ:=((1+h)(x+y)2+h2sx2yhsy2(x+y)2(1+hx)2hsx2(x+y)2(1+hx)y2(x+y)21hδr+x2(x+y)2),(20) where (21) {f1:=x(1+h)1+hxhsxy(x+y)(1+hx),f2:=y+hδy(r+xx+y).(21) Hereafter, we will give local dynamics for the equilibria φ1,2 by stability theory [Citation15–17]. So for φ1, (Equation20) becomes (22) V|φ1:=(11+hhs1+h01hδr+),(22) with (23) λ1=11+h,λ2=1hδr+.(23)

Theorem 2.1

φ1 of model (Equation13) is

  1. a sink if h>2δ(r1);

  2. never source;

  3. a saddle if 0<h<2δ(r1);

  4. non-hyperbolic if h=2δ(r1).

Proof.

By stability theory, φ1 of discrete model (Equation13) is a sink if |λ1,2|<1. Therefore, from (Equation23) if |λ1|=11+h<1 and |λ2|=|1hδr+|<1, that is, h>2δ(r1) then φ1 is a sink. Similarly, it is easy to obtain that φ1 is never source, saddle if 0<h<2δ(r1), and non-hyperbolic if h=2δ(r1).

Now in the following, one has the following result regarding bifurcation set at φ1 if h=2δ(r1) holds.

Theorem 2.2

For the model's semitrivial fixed point φ1, the flip bifurcation set is written as F|φ1:={(h,s,δ,r),h=2δ(r1)}.

Proof.

Since φ1 is non-hyperbolic if h=2δ(r1). Therefore, if h=2δ(r1) holds then λ2|h=2δ(r1)=1 but λ1|h=2δ(r1)=δ(r1)δ(r1)+21or1 which implies that at φ1 eigenvalues criterion for the existence of flip bifurcation holds if (h,s,δ,r) passes F|φ1:={(h,s,δ,r),h=2δ(r1)}.

Now for φ2, (Equation20) becomes (24) V|φ2:=(1hrs(r1)1+h+hs(r1)hr2s1+h+hs(r1)(r1)21+hrδ(r1)),(24) with the following characteristic equation:

(25) det(1hrs(r1)1+h+hs(r1)λhr2s1+h+hs(r1)(r1)21+hrδ(r1)λ)=0,(25)

that is, (26) λ2Λ1λ+Λ2=0,(26) where (27) Λ1=1hrs(r1)1+h+hs(r1)+1+hrδ(r1),Λ2=(1hrs(r1)1+h+hs(r1))(1+hrδ(r1))+h2r2(r1)21+h+hs(r1).(27) From (Equation26), one has (28) λ1,2=Λ1±Δ2,(28) where (29) Δ=Λ124Λ2,=(1hrs(r1)1+h+hs(r1)+1+hrδ(r1))24((1hrs(r1)1+h+hs(r1))(1+hrδ(r1))+h2r2(r1)21+h+hs(r1)).(29)

Theorem 2.3

If Δ<0, then the model's interior equilibrium point φ2 is

  1. a stable focus if 0<δ<1s+sr2r2r with r>max{±s1s,1};

  2. an unstable focus if δ>1s+sr2r2r with r>max{±s1s,1};

  3. non-hyperbolic if δ=1s+sr2r2r.

Proof.

If Δ<0 then from (Equation28) one gets |λ1,2|=(1hrs(r1)1+h+hs(r1))(1+hrδ(r1))+h2r2(r1)21+h+hs(r1)<1. This implies that if 0<δ<1s+sr2r2r with r>max{±s1s,1} then φ2 is a stable focus. Similarly, easy calculation yields the fact that it an unstable focus if δ>1s+sr2r2r with r>max{±s1s,1}, and non-hyperbolic if δ=1s+sr2r2r.

Theorem 2.4

If Δ0 then the model's interior equilibrium point φ2 is

  1. a stable node if δ>42h(1s(r1)2)hr(r1)(2+h+hs(r1)) with s>max{2+hh(r1)2,2+hh(1r)};

  2. an unstable node if 0<δ<42h(1s(r1)2)hr(r1)(2+h+hs(r1)) and s>max{2+hh(r1)2,2+hh(1r)};

  3. non-hyperbolic if δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1)).

Proof.

It is noted that φ2 is a stable node if real roots of (Equation28) satisfying |λ1,2|<1, that is, if δ>42h(1s(r1)2)hr(r1)(2+h+hs(r1)) with s>max{2+hh(r1)2,2+hh(1r)} then φ2 is a stable node. Similarly, easy calculation also implies that it is an unstable node if 0<δ<42h(1s(r1)2)hr(r1)(2+h+hs(r1)), and non-hyperbolic if δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1)).

Now in the following, one has the following result regarding bifurcation set at φ2 if δ=1s+sr2r2r and δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1)) hold, respectively.

Theorem 2.5

For the model's interior equilibrium point φ2, the Neimark–Sacker and flip bifurcation sets, respectively are

  1. N|φ2:={(h,s,δ,r),δ=1s+sr2r2r};

  2. F|φ2:={(h,s,δ,r),δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1))}.

Proof.

(i). It is noted that if δ=1s+sr2r2r holds, then V|φ2 has complex eigenvalues with |λ1,2|δ=1s+sr2r2r=1 which implies that at φ2 eigenvalues criterion for the existence of N-S bifurcation holds if (h,s,δ,r) passes N|φ2:={(h,s,δ,r),δ=1s+sr2r2r}.

(ii). On the other hand, if δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1)) holds then one has λ1|δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1))=1 but λ2|δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1))=2+h(1+s(r1))(1+hrs(r1))(1+h+hs(r1))(2+h+hs(r1))1or1 which implies that at φ2 eigenvalues criterion for the existence of flip bifurcation holds if (h,s,δ,r) passes through the curve F|φ2:={(h,s,δ,r),δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1))}.

3. Analysis of a flip bifurcation for semitrivial fixed point

Based on the local dynamic study for semitrivial fixed point φ1 in Section 2, here we investigate flip bifurcation by bifurcation theory [Citation18–27].

Theorem 3.1

If (h,s,δ,r)F|φ1 then for φ1 the flip bifurcation does not take place in the discrete-time model (Equation13).

Proof.

Since model (Equation13) is invariant under y = 0, and so one restrict it on the line y = 0 where it becomes (30) xt+1=xt(1+h)1+hxt.(30) From (Equation30), one has (31) f(h,x):=x(1+h)1+hx.(31) Finally, if x=x=1 and h=h=2δ(r1), from (Equation31), one has (32) fx|h=2δ(r1),x=1:=δ(r1)δ(r1)+21,(32) (33) 2fx2|h2δ(r1),x=1:=4δ(r1)(δ(r1)+2)20,(33) and (34) fh|h=2δ(r1),x=1:=0.(34) From (Equation32) and (Equation34), one can be concluded that the flip bifurcation does not take place if (h,s,δ,r)F|φ1.

4. Analysis of a Neimark–Sacker bifurcation for interior fixed point

Based on the local dynamic study for the equilibrium φ2 in Section 2, we investigate the Neimark–Sacker bifurcation if (h,s,δ,r)N|φ2 where δ is regarded as a bifurcation parameter.

Theorem 4.1

At φ2, discrete model (Equation13) undergoes N–S bifurcation if (h,s,δ,r)N|φ2.

Proof.

Recall that if δ is a bifurcation parameter then perturbation system of discrete model (Equation13) takes the form (35) {xt+1=xt(1+h)1+hxthsxtyt(xt+yt)(1+hxt),yt+1=yt+h(δ+ϵ)yt(r+xtxt+yt),(35) where δ=δ+ϵ and 0<|δ|1. For the perturbation system (Equation35), the roots of V|φ2 are (36) λ1,2:=Λ1(ϵ)±ι4Λ2(ϵ)Λ12(ϵ)2,(36) where Λ1(ϵ)=1hrs(r1)1+h+hs(r1)+1+hr(δ+ϵ)(r1) and Λ2(ϵ)=(1hrs(r1)1+h+hs(r1))(1+hr(δ+ϵ)(r1))+h2r2s(δ+ϵ)(r1)21+h+hs(r1). Now due to the fact that (h,s,δ,r)N|φ2, the characteristic equation (Equation36) of V|φ2 has two conjugate complex root with |λ1,2|=(1hrs(r1)1+h+hs(r1))(1+hr(δ+ϵ)(r1))+h2r2s(δ+ϵ)(r1)21+h+hs(r1), and subsequently one gets the following quantity by noticing that δ=1s+sr2r2r: (37) d|λ1,2||ϵ=0=hr(r1)2(1+h+hs(r1))0.(37) Moreover for ϵ=0, the characteristics roots of (Equation36) must not be found at the intersections of unit circle with coordinate axes. Therefore, one need to suppose that λ11 and λ21(=1,2,3,4) for ϵ=0, which is equivalent to Λ1(0)=1hrs(r1)1+h+hs(r1)+1+hrδ(r1)2,0,1,2. But if δ=1s+sr2r2r holds, then Λ2(0)=1, and so, Λ1(0)2,2. Therefore, Λ1(0)=1hrs(r1)1+h+hs(r1)+1+hrδ(r1)0,1, and so, by straightforward calculation one obtains (38) s2h(2+r)±48r+hr(4+hr)2h(r21),1h(2+r)±34r+hr(2+hr)2h(r21).(38) Now to transform φ2 one use the following transformations: (39) ut=xtx,vt=yty,(39) with x=1s+sr and y=(1r)(1s+sr)r. From (Equation39) and (Equation35), we get (40) {ut+1=(ut+x)(1+h)1+h(ut+x)hs(ut+x)(vt+y)((ut+x)+(vt+y))(1+h(ut+x))x,vt+1=(vt+y)+h(δ+ϵ)(vt+y)(r+(ut+x)(ut+x)+(vt+y))y.(40) If ϵ=0, then we will derive normal form of (Equation40). On expanding (Equation40) by Taylor series up to second-order at origin, one gets (41) {ut+1=Ω111ut+Ω121vt+Ω131ut2+Ω141utvt+Ω151vt2,vt+1=Ω112ut+Ω122vt+Ω132ut2+Ω142utvt+Ω152vt2,(41)

where its coefficients are mentioned in (EquationA1). Now (Equation41) takes the form: (42) {xt+1=ηxtζyt+P¯(xt,yt),yt+1=ζxt+ηyt+Q¯(xt,yt),(42) by following matrix transformation (43) (utvt):=(Ω1210ηΩ111ζ)(xtyt),(43) where (44) {P¯(xt,yt)=r11xt2+r12xtyt+r13yt2,Q¯(xt,yt)=r21xt2+r22xtyt+r23yt2,(44) and the quantities η, ζ and rij(i,j=1,2,3) are depicted in (EquationA2). Furthermore, from (Equation44) we get 2P¯xt2|φ2=(0,0)=2r11, 2P¯xtyt|φ2=(0,0)=r12, 2P¯yt2|φ2=(0,0)=2r13, 3P¯xt3|φ2=(0,0)=0, 3P¯xt2yt|φ2=(0,0)=0, 3P¯xtyt2|φ2=(0,0)=0, 3P¯yt3|φ2=(0,0)=0, 2Q¯xt2|φ2=(0,0)=2r21, 2Q¯xtyt|φ2=(0,0)=r22, 2Q¯yt2|φ2=(0,0)=2r23, 3Q¯xt3|φ2=(0,0)=3Q¯xt2yt|φ2=(0,0)=3Q¯xtyt2|φ2=(0,0)=3Q¯yt3|φ2=(0,0)=0. Now the following discriminatory quantity should not be zero in order to guarantee the occurrence of Neimark–Sacker bifurcation at φ2 of system (Equation42): (45) :=((12λ¯)λ¯21λϱ11ϱ20)12ϱ112ϱ022+(λ¯ϱ21),(45) where (46) ϱ02=18(2P¯xt22P¯yt2+22Q¯xtyt+ι(2Q¯xt22Q¯yt2+22P¯xtyt))|φ2=(0,0),ϱ11=14(2P¯xt2+2P¯yt2+ι(2Q¯xt2+2Q¯yt2))|φ2=(0,0),ϱ20=18(2P¯xt22P¯yt2+22Q¯xtyt+ι(2Q¯xt22Q¯yt222P¯xtyt))|φ2=(0,0),ϱ21=116(3P¯xt3+3P¯yt3+3Q¯xt2yt+3Q¯yt3+ι(3Q¯xt3+3Q¯xtyt23P¯xt2yt3P¯yt3))|φ2=(0,0).(46) Using partial derivatives, (Equation46) becomes (47) ϱ02=14(r11r13+r22+ι(r21r23+r12)),ϱ11=12(r11+r13+ι(r21+r23)),ϱ20=14(r11r13+r22+ι(r21r23r12)),ϱ21=0.(47) From (Equation45) the condition 0 holds as (h,s,δ,r)N|φ2 then Neimark–Sacker bifurcation takes place, and additionally supercritical (subcritical) Neimark–Sacker bifurcation take place if <0 (>0).

5. Analysis of a flip bifurcation for interior fixed point

Based on the local dynamic study for the equilibrium φ2 in Section 2, we investigate the flip bifurcation if (h,s,δ,r)F|φ2 where δ is regarded as a bifurcation parameter.

Theorem 5.1

At φ2, discrete model (Equation13) undergoes flip bifurcation if (h,s,δ,r)F|φ2.

Proof.

Recall that if δ is a bifurcation parameter then the discrete model (Equation13) takes the form (Equation35), and by (Equation39) it further becomes (48) ut+1=Ω111ut+Ω121vt+Ω131ut2+Ω141utvt+Ω151vt2,vt+1=Ω112ut+Ω122vt+Ω132ut2+Ω142utvt+Ω152vt2+k1utϵ+k2vtϵ+k3ut2ϵ+k4utvtϵ+k5vt2ϵ,(48) where it coefficients are depicted in (EquationA3). Now using following transformation: (49) (utvt):=(12hs(1+r)2+hhsr212(1r)(1+h(1+srs))r(2+hhs+hsr))(xtyt),(49) Equation (Equation48) gives (50) (xt+1yt+1)=(100λ2)(xtyt)+(P^(ut,vt,ϵ)Q^(ut,vt,ϵ)),(50) where (51) P^={2hrs(r1)(1+h+hsrhs)4+h(1+s(r1))(4+h+hsr2hs)(Ω131ut2+Ω141utvt+Ω151vt2)+hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))×(Ω132ut2+Ω142utvt+Ω152vt2+k1ϵut+k2ϵvt+k3ϵut2+k4ϵutvt+k5ϵvt2)},Q^={(2+h(1+srs))(2+h(1+s(r1)2))4+h(1+s(r1))(4+h+hs(r21))(Ω131ut2+Ω141utvt+Ω151vt2)hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))×(Ω132ut2+Ω142utvt+Ω152vt2+k1ϵut+k2ϵvt+k3ϵut2+k4ϵutvt+k5ϵvt2)},(51)

Now we calculate the centre manifold FCO at O in a small neighborhood of ϵ=0 where mathematically it can be expressed as (53) FCO={(xt,yt):yt=v0ϵ+v1xt2+v2ϵxt+v3ϵ3+O((|xt|+|ϵ|)3)},(53) with (54) {v1={11λ2((2+h(1+srs))(2+h(1+s(r1)2))4+h(1+s(r1))(4+h+hs(r21))×(Ω131+2hs(1+r)2+hhsr2Ω141+(2hs(1+r)2+hhsr2)2Ω151)hr2s(2+h(1+srs))4+h(1+s(r1))(4+h+hs(r21))×(Ω132+2hs(1+r)2+hhsr2Ω142+(2hs(1+h)2+hhsr2)2Ω152))},v2={11λ2(hr2s(2+h(1+srs))4+h(1+s(r1))(4+h+hs(r21))×(k1+2hs(1+r)2+hhsr2k2))},v0=v3=0.(54) Finally, one write (Equation50) restrict to FcO as (55) f1(xt)=xt+m1xt2+m2xtϵ+m3xt2ϵ+m4xtϵ2+m5xt3+O((|xt|+|ϵ|)4),(55) where

(56) {m1={2hrs(r1)(1+h+hs(r1))4+h(1+s(r1))(4+h+hs(r21))×(Ω131+(2hs(1+r)2+hhsr2)Ω141(2hs(1+r)2+hhsr2)2Ω151)+hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))×(Ω132+(2hs(1+r)2+hhsr2)Ω142(2hs(1+r)2+hhsr2)2Ω152)},m2=hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))(k1+2hs(1+r)2+hhsr2k2),m3={2hrs(r1)(1+h+hs(r1))4+h(1+s(r1))(4+h+hs(r21))×(2Ω131v2+Ω141v2(2(1r)(1+h(1+srs))r(2+hhs+hsr)+2hs(1+r)2+hhsr2)+Ω151v2(2(2hs(1+r)2+hhsr2)(2(1r)(1+h(1+srs))r(2+hhs+hsr))))+hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))(2Ω132v2+Ω142v2(2(1r)(1+h(1+srs))r(2+hhs+hsr)+2hs(1+r)2+hhsr2)+Ω152v2(2(2hs(1+r)2+hhsr2)(2(1r)(1+h(1+srs))r(2+hhs+hsr)))+k1v1+(2(1r)(1+h(1+srs))r(2+hhs+hsr))k2v1+k3+(2hs(1+r)2+hhsr2)k4+(2hs(1+r)2+hhsr2)2k5)},m4={hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))(k1v2+2(1r)(1+h(1+srs))r(2+hhs+hsr)k2v2)},m5={2hrs(r1)(1+h+hs(r1))4+h(1+s(r1))(4+h+hs(r21))×(2Ω131v1+Ω141v1(2(1r)(1+h(1+srs))r(2+hhs+hsr)+2hs(1+r)2+hhsr2)+Ω151v1(2(2hs(1+r)2+hhsr2)(2(1r)(1+h(1+srs))r(2+hhs+hsr))))+hr2s(2+h(1+srs))4+h(1+s(1+r))(4+h+hs(r21))(2Ω132v12(1r)(1+h(1+srs))r(2+hhs+hsr)+Ω142v1(2(1r)(1+h(1+srs))r(2+hhs+hsr)+2hs(1+r)2+hhsr2)+Ω152v1(2(2hs(1+r)2+hhsr2)(2(1r)(1+h(1+srs))r(2+hhs+hsr))))}.(56)

Finally, in order for the existence of flip bifurcation at φ2, the following two discriminatory quantities 1 and 2 should not be zero [Citation18, Citation19]: (57) 1:=(2f1xtϵ+12f1ϵ2f1xt2)|(0,0)=m2,2:=(163f1xt3+(122f1xt2)2)|(0,0)=m5+m12.(57) Consequently, based on above calculation, we can say that if (h,s,δ,r)F|φ2 and 20 then at φ2 flip bifurcation takes place. Besides, the period-2 points bifurcate from φ2 of model (Equation13) are stable (unstable) if 2>0 (2<0).

6. Chaos control

In the present section, we will apply the following two feedback control strategies to the model (Equation13) to get a stable trajectory:

6.1. By OGY method

In order to study chaos, we fist use OGY method, which was proposed by Ott et al. [Citation28], for the discrete model (Equation13). By existing chaos theory [Citation29, Citation30], one can write the model (Equation13) as follows: (58) {xt+1=xt(1+h)1+hxthsxtyt(xt+yt)(1+hxt)=f1(xt,yt,s),yt+1=yt+yt(r+xtxt+yt)=f2(xt,yt,s),(58) where s represents the control parameter under which one can acquire the desired chaos control by small perturbations. To do this, one restrict it where s(s0σ,s0+σ) with σ>0, and s0 indicates nominal value corresponds to chaotic region. The trajectory is moved in the direction of the target orbit using the stabilizing feedback control strategy. Assuming that φ2 be unstable equilibrium point of the model (Equation13), which is in the chaotic region created by the occurrence of N-S bifurcation, then the given below linear map can be used to approximate the model (Equation58): (59) (xt+1xyt+1y)V(x,y,s0)(xtxyty)+P(ss0),(59) where (60) V|(x,y,s0)=(f1(x,y,s0)xf1(x,y,s0)yf2(x,y,s0)xf2(x,y,s0)y)=(1hrs0(r1)1+h+hs0(r1)hr2s01+h+hs0(r1)(r1)21+hrδ(r1)),(60) and (61) P=(f1(x,y,s0)sf2(x,y,s0)s)=(h(r1)(1s0+s0r)1+h+hs0(r1)0).(61) Moreover, the model (Equation58) is controlled provided that the following matrix (62) C=(P:VP)=(h(1+r)(1s0+s0r)1+h+hs0(1+r)0h(1+r)(1hrs0(1+r))(1s0+s0r)(1+h+hs0(1+r))2h2δ(1+r)3(1s0+s0r)1+h+hs0(1+r)),(62) is of rank 2. Additionally, it must be understood that δ cannot be used in this situation as a control parameter to utilize the OGY method to the model (Equation13). As, if we take f1(xt,yt,δ)=xt(1+h)1+hxthsxtyt(xt+yt)(1+hxt) and f2(xt,yt,δ)=yt+yt(r+xtxt+yt), then Rank(C) becomes zero. So, to apply the OGY method to the model (Equation13), s is taken as a control parameter. Now, if one take ss0=T(xtxyty), where T=(k1k2), then the model (Equation59) can be written as (63) (xt+1xyt+1y)(VPT)(xtxyty).(63) Furthermore, the corresponding controlled model of (Equation13) is (64) {xt+1=xt(1+h)1+hxth(s0k1(xtx)k2(yty))xtyt(xt+yt)(1+hxt),yt+1=yt+yt(r+xtxt+yt).(64) By stability theory φ2 is a sink iff roots of VPT satisfying |λ1,2|<1 where its variational matrix is (65) VPT=(1hrs0(r1)1+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k1(r1)2hr2s01+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k21+hrδ(r1)).(65) The characteristic equation of VPT is (66) {λ2(1hrs0(r1)1+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k1+1+hrδ(r1))λ+(1hrs0(r1)1+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k1)(1+hrδ(r1))+(r1)2(hr2s01+h+hs0(r1)+h(r1)(1s0+s0r)1+h+hs0(r1)k2)=0.(66) If λ1,2 denotes the roots of (Equation66) then (67) λ1+λ2=1hrs0(r1)1+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k1+1+hrδ(r1),(67) and (68) λ1λ2=(1hrs0(r1)1+h+hs0(r1)h(r1)(1s0+s0r)1+h+hs0(r1)k1)(1+hrδ(r1))+(r1)2(hr2s01+h+hs0(r1)+h(r1)(1s0+s0r)1+h+hs0(r1)k2).(68) Next we take λ1=±1 and λ1λ2=1 to get lines of marginal stability for (Equation64). Also, these restriction give the fact that characteristics roots satisfying |λ1,2|<1. Assuming that λ1λ2=1 then from (Equation68) one has (69) L1:h(r1)(1s0+s0r)(1+hrδ(r1))k1h2δ(r1)3(1s0+s0r)k2+hhrδ(r1)+hs0(r21)=0.(69) If λ1=1 then from (Equation67) and (Equation68), one gets (70) L2:h2(r1)2(1s0+s0r)k1h2δ(r1)3(1s0+s0r)k2+h2(r1)(1s0+s0r))=0.(70) Finally, if λ1=1 then from (Equation67) and (Equation68), one gets (71) L3:h(r1)(1s0+s0r)(2+hrδ(r1))k1h2δ(r1)3(1s0+s0r)k242h+2hs0(r1)2hrδ(1+r)(2+h(1+s0rs0))=0.(71) Then, stable eigenvalues lie within the triangular region (k1k2)-plane bounded by the straight lines L1,2,3 for parametric values h, s, r and δ.

6.2. By hybrid control feedback

We use a hybrid control feedback method to control the chaos due to the emergence of flip bifurcation in the model (Equation13) by existing theory [Citation31]. If the model (Equation13) undergoes bifurcation at φ2 then one can write corresponding controlled model as (72) {xt+1=α(xt(1+h)1+hxthsxtyt(xt+yt)(1+hxt))+(1α)xt,yt+1=α(yt+yt(r+xtxt+yt))+(1α)yt,(72) where 0<α<1 and controlled strategy (Equation72) is a combination of feedback control as well as parameter perturbation. The V|φ2 is (73) V|φ2=(1+αh(1+ssr2)1+h+hs(r1)αhr2s1+h+hs(r1)αhδ(r1)21+αhrδ(r1)).(73) The characteristics equation of V|φ2 evaluated at φ2 is (74) λ2Λ1λ+Λ2=0(74) where (75) Λ1=2+αh(1+ssr2)1+h+hs(r1)+αhδr(r1),Λ2=(1+αh(1+ssr2)1+h(1+s(r1)))(1+αhrδ(r1))+α2h2r2(r1)21+h(1+hs(r1)).(75) So, based on stability theory one has the result:

Lemma 1

Equilibrium φ2 of model (Equation72) is a sink iff |2+αh(1+ssr2)1+h+hs(r1)+αhδr(r1)|<1+(1+αh(1+ssr2)1+h+hs(r1))(1+αhrδ(r1))+α2h2r2(r1)21+h+hs(r1)<2.

Proof.

The necessary and sufficient condition for roots of (Equation74) satisfying |λ1,2| imply that φ2 of model (Equation72) is a sink iff |2+αh(1+ssr2)1+h+hs(r1)+αhδr(r1)|<1+(1+αh(1+ssr2)1+h+hs(r1))(1+αhrδ(r1))+α2h2r2(r1)21+h+hs(r1)<2.

7. Numerical simulations

Example 7.1

If h = 0.6, s = 1.37, r = 0.5, δ=[0.1,1.1] with (x0,y0)=(0.03,0.04) then at δ=0.11000000000000032 discrete model (Equation13) undergoes the N-S bifurcation. The bifurcation diagrams and Maximum Lyapunov exponent are drawn in Figure . Further, at (h,s,r,δ)=(0.6,1.37,0.5,0.11000000000000032) model (Equation13) has interior equilibrium point φ2=(0.31499999999999995,0.31499999999999995) and moreover from (Equation24) one gets: (76) V|φ2=(0.31499999999999995,0.31499999999999995)=(1.01387720773759440.0165000000000000460.172834314550042080.9834999999999999),(76) with (77) λ21.9973772077375944λ+1=0.(77) The roots of (Equation77) are λ1,2=0.9986886038687972±0.05119641103234161ι with |λ1,2|=1, and so from (i) of Theorem 2.5 one has the Neimark–Sacker bifurcation set (h,s,r,δ)=(0.6,1.37,0.5,0.11000000000000032)N|φ2=(0.31499999999999995,0.31499999999999995). Furthermore, by fixing h = 0.6, s = 1.37, r = 0.5, and varying the value of bifurcation parameter δ<0.11000000000000032 then one can obtain that respective interior equilibrium solution is a stable focus. To verify this if δ=0.10000<0.11000000000000032 then Figure (a) indicates that φ2=(0.31499999999999995,0.31499999999999995) is a stable focus. Similarly if one varies δ=0.10299, 0.10500, 0.10880, 0.10987, 0.10999<0.11000000000000032 then Figure (b–f) also show that, for each variation of δ, the equilibrium solution φ2=(0.31499999999999995,0.31499999999999995) is also stable focus. On the other hand, if one varies the bifurcation parameter δ>0.11000000000000032 then we can concluded that respective interior equilibrium solution is an unstable focus, and as a result supercritical Neimark–Sacker bifurcation must take place. For instance, if h = 0.6, s = 1.37, r = 0.5 then from (Equation37) the non-degenerate conditions, i.e. d|λ1,2||ϵ=0=0.063078216989066430 holds, and moreover if δ=0.11000000000000032 then from (EquationA1) and (EquationA2) one gets: (78) {Ω111=1.0138772077375944,Ω121=0.17283431455004208,Ω131=0.2372883502540626,Ω141=0.46146372048017426,Ω151=0.2743401818254637,Ω112=0.016500000000000046,Ω122=0.9834999999999999,Ω132=0.02619047619047627,Ω142=0.05238095238095254,Ω152=0.02619047619047627,(78) and (79) {η=0.9986886038687973,ζ=0.05119641103234052.(79) Utilizing (Equation78) and (Equation79) into (EquationA2), one gets: (80) {r11=3.2139366880763243,r12=0.021156708511249955,r13=0.0041604325437973955,r21=0.0014241879052949461,r22=0.0634844121741236,r23=0.0025751472836634173.(80) Now in view of (Equation80) and (Equation47), one obtains (81) {ϱ02=0.7886531771114995+0.004289343330572898ι,ϱ11=1.6048881277662634+0.0005754796891842356ι,ϱ20=0.78865317711149950.0062890109250520795ι,ϱ21=0.(81) Now finally, using (Equation81) along with λ,λ¯=0.9986886038687973±0.05119641103234052ι into (Equation45) one gets: =3.9314493377802>0 which give the fact that closed invariant curve must exists, and hence discrete model undergoes subcritical Neimark–Sacker at indicated interior fixed point φ2=(0.31499999999999995,0.31499999999999995) (see Figure (a)). In similar manner, we can also obtain that if one varies δ=0.125670,0.130000,0.139999,0.141100,0.149999,0.158896,0.161231,0.167890>0.11000000000000032 then stable invariant curves also appears which are depicted in Figure (b–i). Therefore, in conclusion we can say that numerical simulations in Example 7.1 agree with theoretical results obtained in Theorem 2.3, (i) of Theorem 2.5 and Theorem 4.1.

Figure 1. Bifurcation diagrams (B.Ds) with MLE of discrete model (Equation13). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) B.D for s and xt. (e) B.D for s and yt. (f) MLEs.

Figure 1. Bifurcation diagrams (B.Ds) with MLE of discrete model (Equation13(13) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt),yt+1=yt+hδyt(−r+xtxt+yt),(13) ). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) B.D for s and xt. (e) B.D for s and yt. (f) MLEs.

Figure 2. Phase portrait of discrete model (Equation13) with (0.03,0.04). (a) If δ=0.10000. (b) If δ=0.10299. (c) If δ=0.10500. (d) If δ=0.10880. (e) If δ=0.10987. (f) If δ=0.10999.

Figure 2. Phase portrait of discrete model (Equation13(13) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt),yt+1=yt+hδyt(−r+xtxt+yt),(13) ) with (0.03,0.04). (a) If δ=0.10000. (b) If δ=0.10299. (c) If δ=0.10500. (d) If δ=0.10880. (e) If δ=0.10987. (f) If δ=0.10999.

Figure 3. Invariant closed curves of discrete model (Equation13) with (0.03,0.04). (a) If δ=0.110999. (b) If δ=0.125670. (c) If δ=0.130000. (d) If δ=0.139999. (e) If δ=0.141100. (f) If δ=0.149999. (g) If δ=0.158896. (h) If δ=0.161231. (i) If δ=0.167890.

Figure 3. Invariant closed curves of discrete model (Equation13(13) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt),yt+1=yt+hδyt(−r+xtxt+yt),(13) ) with (0.03,0.04). (a) If δ=0.110999. (b) If δ=0.125670. (c) If δ=0.130000. (d) If δ=0.139999. (e) If δ=0.141100. (f) If δ=0.149999. (g) If δ=0.158896. (h) If δ=0.161231. (i) If δ=0.167890.

Example 7.2

If h = 1.28, s = 2.5, r = 0.798, δ=[10.1,16.99] with (x0,y0)=(0.495,0.125301) then at δ=11.591710484782386 discrete model (Equation13) undergoes the flip bifurcation. The flip bifurcation diagrams and Maximum Lyapunov exponent are drawn in Figure . Further at (h,s,r,δ)=(1.28,2.5,0.798,11.591710484782386) discrete model (Equation13) has equilibrium solution φ2=(0.4950000000000001,0.12530075187969925) and moreover from (Equation24) one gets: (82) V|φ2=(0.92790597453476960.60542483791495721.24741234084231151.3917278250303777),(82) with λ1=1 and λ2=0.53617814950439311or1, and hence based on these simulations one can obtain that for parametric values (h,s,r,δ)=(1.28,2.5,0.798,11.591710484782386)F|φ2=(0.4950000000000001,0.12530075187969925). Moreover in this parametric domain, from (EquationA3), (Equation54) and (Equation56) one gets: (83) {Ω111=0.9279059745347696,Ω121=1.2474123408423115,Ω131=0.5982005795778922,Ω141=0.040685691442227376,Ω151=2.0109798949336652,Ω112=0.6054248379149572,Ω122=1.3917278250303777,Ω132=0.976018223547749,Ω142=7.711510320703999,Ω152=15.232141673073743,k1=0.05222911999999998,k2=0.20633088,k3=0.0841996722424242,k4=0.6652607767272725,k5=1.3140547025454543,(83) (84) {v0=0,v1=12.929077600120031,v2=0.10826715884189564,v3=0,(84) and (85) {m1=21.71571925204092,m2=0.2165343176837912,m3=2.143746543978618,m4=0.0011046972084443997,m5=27.41483556845985.(85) Utilizing (Equation85) into (Equation57) one gets: 1=0.21653431768379120 and 2=444.1576270650008>0. Since 2=444.1576270650008>0 and so it can be concluded that stable period-2 points bifurcate from φ2=(0.4950000000000001,0.12530075187969925). Therefore, in conclusion we can say that numerical simulations in Example 7.2 agree with theoretical results obtained in Theorem 2.4, (ii) of Theorem 2.5 and Theorem 5.1.

Figure 4. Flip B.Ds with MLE of discrete model (Equation13). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) B.D for h and xt. (e) B.D for h and yt. (f) B.D for s and xt. (g) B.D for s and yt. (h) MLEs.

Figure 4. Flip B.Ds with MLE of discrete model (Equation13(13) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt),yt+1=yt+hδyt(−r+xtxt+yt),(13) ). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) B.D for h and xt. (e) B.D for h and yt. (f) B.D for s and xt. (g) B.D for s and yt. (h) MLEs.

Example 7.3

If h = 0.5, r = 0.4, δ=0.38333333333333336, s[0.1,1.33] with (x0,y0)=(0.02,0.03) then model (Equation13) undergoes Neimark–Sacker bifurcation. The maximum Lyapunov exponents with bifurcation diagrams are drawn in Figure . Further if s = 1.3 then model (Equation13) has equilibrium φ2=(0.21999999999999997,0.32999999999999996) where roots of characteristics equations of V|φ2=(0.21999999999999997,0.32999999999999996) are λ1,2=0.9977207207207206±0.06747861471996602ι with |λ1,2|=1. In order to apply OGY control method for model (Equation13), if s0=0.3 then it has equilibrium φ2=(0.82,1.2299999999999998). So, model (Equation64) becomes (86) {xt+1=xt+0.5xt(1xt)0.5(0.3k1(xt0.82)k2(yt1.2299999999999998))xtytxt+yt,yt+1=yt+0.5(0.38333333333333336)yt(0.4+xtxt+yt),(86) where its variation matrix VBK is (87) VBK=(0.7347517730496455+0.17446808510638295k10.06899940.017021276595744685+0.17446808510638295k10.9540004).(87) Now for marginal stability lines L1,L2 and L3 are (88) L1:3.340151253503999k1+16.523026880255998k2+13.10398752=0,(88) (89) L2:11.015351253503999k1+16.523026880255998k218.358918755839998=0,(89) and (90) L3:4.335048746496k1+16.523026880255998k210.60110620416=0.(90) So, lines (Equation88), (Equation89) and (Equation90) determine triangular region that gives |λ1,2|<1(see Figure ).

Figure 5. B.Ds with MLE of discrete model (Equation13). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) MLEs.

Figure 5. B.Ds with MLE of discrete model (Equation13(13) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt),yt+1=yt+hδyt(−r+xtxt+yt),(13) ). (a) B.D for xt. (b) B.D for yt. (c) B.D for xt and yt. (d) MLEs.

Figure 6. Region of stability where |λ1,2|<1.

Figure 6. Region of stability where |λ1,2|<1.

Example 7.4

Finally, if h=1.28,s=2.5,r=0.798,δ=11.591710484782386 with (x0,y0)=(1.3,1.12) then model (Equation13) undergoes flip bifurcation. For this, by applying hybrid strategy to get stable orbit at φ2=(0.4950000000000001,0.12530075187969925). For this model (Equation72) takes the form (91) {xt+1=α(xt+1.28xt(1xt)(1.28)(2.5)xtytxt+yt)+(1α)xt,yt+1=α(yt+(1.28)(11.591710484782386)yt(0.798+xtxt+yt))+(1α)yt,(91) where Vφ2=(0.4950000000000001,0.12530075187969925) is (92) Vφ2=(10.07209402546523036α0.6054248379149572α1.2474123408423115α12.3917278250303777α),(92) with (93) λ2(22.463821850495606α)λ+12.463821850495606α+0.9276437009912135α2=0.(93) Furthermore, roots of (Equation93) satisfying |λ1,2|<1 if 0<α<1. So, for the allowed interval of control parameter α the flip bifurcation is completely eliminated. If α=0.999 then for controlled model (Equation91), plots of t vs. xt and yt are drawn in Figure .

Figure 7. Plot of t vs. xt and yt for controlled system (Equation58).

Figure 7. Plot of t vs. xt and yt for controlled system (Equation58(58) {xt+1=xt(1+h)1+hxt−hsxtyt(xt+yt)(1+hxt)=f1(xt,yt,s),yt+1=yt+hδyt(−r+xtxt+yt)=f2(xt,yt,s),(58) ).

8. Conclusion

This work explored the local dynamics at fixed points, chaos, bifurcations of a discrete prey–predator model (Equation13) in the region: R+2={(x,y):x,y0}. We studied local dynamics at semitrivial and interior fixed points φ1,2 of model (Equation13), and explored that φ1 of model (Equation13) is a sink if h>2δ(r1), never source, saddle if 0<h<2δ(r1), and non-hyperbolic if h=2δ(r1). Further it proved that φ2 of model (Equation13) is a stable focus if 0<δ<1s+sr2r2r with r>max{±s1s,1}, an unstable focus if δ>1s+sr2r2r with r>max{±s1s,1}, non-hyperbolic condition if δ=1s+sr2r2r, stable node if δ>42h(1s(r1)2)hr(r1)(2+h+hs(r1)) with s>max{2+hh(r1)2,2+hh(1r)}, an unstable node if 0<δ<42h(1s(r1)2)hr(r1)(2+h+hs(r1)) with s>max{2+hh(r1)2,2+hh(1r)}, and non-hyperbolic if δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1)). In order to studied the bifurcation analysis at φ1,2, we first identified bifurcation sets for understudied model (i) flip bifurcation set F|φ1:={(h,s,δ,r),h=2δ(r1)} at φ1, (ii) Neimark–Sacker bifurcation set N|φ2:={(h,s,δ,r),δ=1s+sr2r2r} at φ2 (iii) flip bifurcation set whose mathematical expression is F|φ2:={(h,s,δ,r),δ=42h(1s(r1)2)hr(r1)(2+h+hs(r1))} at φ2, and then proved that at φ1 flip bifurcation did not take place if (h,s,δ,r)F|φ1 but at φ2 model subjected Neimark–Sacker and flip bifurcations if (h,s,δ,r)N|φ2 and (h,s,δ,r)F|φ2, respectively. Furthermore, OGY and Hybrid control strategies are utilized to control chaos in the under study discrete model due to occurrence of Neimark–Sacker and flip bifurcations. Finally numerical simulations depicted to verify theoretical results.

Supplemental material

Source_files.rar

Download ()

Acknowledgments

Authors declare that they got no funding on any part of this research.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Freedman HI. Deterministic mathematical models in population ecology. New York: Marcel Dekker Incorporated; 1980.
  • Arditi R, Ginzburg LR, Akcakaya HR. Variation in plankton densities among lakes: a case for ratio-dependent predation models. Am Nat. 1991;138(5):1287–1296. doi: 10.1086/285286
  • Akcakaya HR, Arditi R, Ginzburg LR. Ratio-dependent predation: an abstraction that works. Ecology. 1995;76(3):995–1004. doi: 10.2307/1939362
  • Cosner C, DeAngelis DL, Ault JS, et al. Effects of spatial grouping on the functional response of predators. Theor Popul Biol. 1999;56(1):65–75. doi: 10.1006/tpbi.1999.1414
  • Arditi R, Ginzburg LR. Coupling in predator–prey dynamics: ratio-dependence. J Theor Biol. 1989;139(3):311–326. doi: 10.1016/S0022-5193(89)80211-5
  • Kuang Y, Beretta E. Global qualitative analysis of a ratio-dependent predator–prey system. J Math Biol. 1998;36(4):389–406. doi: 10.1007/s002850050105
  • Hsu SB, Hwang TW, Kuang Y. Global analysis of the Michaelis–Menten-type ratio-dependent predator–prey system. J Math Biol. 2001;42(6):489–506. doi: 10.1007/s002850100079
  • Singh A, Malik P. Bifurcations in a modified Leslie–Gower predator–prey discrete model with Michaelis–Menten prey harvesting. J Appl Math Comput. 2021;67(1-2):143–174. doi: 10.1007/s12190-020-01491-9
  • Ma R, Bai Y, Wang F. Dynamical behavior analysis of a two-dimensional discrete predator–prey model with prey refuge and fear factor. J Appl Anal Comput. 2020;10(4):1683–1697.
  • Santra PK, Panigoro HS, Mahapatra GS. Complexity of a discrete-time predator–prey model involving prey refuge proportional to predator. Jambura J Math. 2022;4(1):50–63. doi: 10.34312/jjom.v4i1
  • Chen J, Chen Y, Zhu Z, et al. Stability and bifurcation of a discrete predator–prey system with Allee effect and other food resource for the predators. J Appl Math Comput. 2023;69(1):529–548.
  • Tunç C, Tunç O. Qualitative analysis for a variable delay system of differential equations of second order. J Taibah Univ Sci. 2019;13(1):468–477. doi: 10.1080/16583655.2019.1595359
  • Tunç C. Stability to vector Lienard equation with constant deviating argument. Nonlinear Dyn. 2013;73(3):1245–1251. doi: 10.1007/s11071-012-0704-8
  • Tunç C. A note on boundedness of solutions to a class of non-autonomous differential equations of second order. Appl Anal Discret Math. 20104:361–372. doi: 10.2298/AADM100601026T
  • Wikan A. Discrete dynamical systems with an introduction to discrete optimization problems. London (UK): Bookboon. com; 2013.
  • Kulenović MRS, Ladas G. Dynamics of second order rational difference equations: with open problems and conjectures. New York: Chapman and Hall/CRC; 2001.
  • Camouzis E, Ladas G. Dynamics of third-order rational difference equations with open problems and conjectures. New York: CRC Press; 2007.
  • Guckenheimer J, Holmes P. Nonlinear oscillations, dynamical systems, and bifurcations of vector fields. New York: Springer Science & Business Media; 2013.
  • Kuznetsov YA, Kuznetsov IA, Kuznetsov Y. Elements of applied bifurcation theory. New York: Springer; 1998.
  • Rana SM. Chaotic dynamics and control of discrete ratio-dependent predator–prey system. Discrete Dyn Nat Soc. 2017;2017:1–13.
  • Al-Basyouni KS, Khan AQ. Discrete-time predator–prey model with bifurcations and chaos. Math Probl Eng. 2020;2020:1–14. doi: 10.1155/2020/8845926
  • Chakraborty P, Ghosh U, Sarkar S. Stability and bifurcation analysis of a discrete prey–predator model with square-root functional response and optimal harvesting. J Biol Syst. 2020;28(01):91–110. doi: 10.1142/S0218339020500047
  • Liu W, Cai D. Bifurcation, chaos analysis and control in a discrete-time predator–prey system. Adv Differ Equ. 2019;2019(1):1–22. doi: 10.1186/1687-1847-2012-1
  • Agiza HN, Elabbasy EM, El-Metwally H, et al. Chaotic dynamics of a discrete prey–predator model with Holling type II. Nonlinear Anal Real World Appl. 2009;10(1):116–129. doi: 10.1016/j.nonrwa.2007.08.029
  • Liu X, Xiao D. Complex dynamic behaviors of a discrete-time predator–prey system. Chaos Solit Fractals. 2007;32(1):80–94. doi: 10.1016/j.chaos.2005.10.081
  • Khan AQ, Ma J, Xiao D. Bifurcations of a two-dimensional discrete time plant-herbivore system. Commun Nonlinear Sci Numer Simul. 2016;39:185–198. doi: 10.1016/j.cnsns.2016.02.037
  • Khan AQ, Ma J, Xiao D. Global dynamics and bifurcation analysis of a host-parasitoid model with strong Allee effect. J Biol Dyn. 2017;11(1):121–146. doi: 10.1080/17513758.2016.1254287
  • Ott E, Grebogi C, Yorke JA. Controlling chaos. Phys Rev Lett. 1990;64(11):1196–1199. doi: 10.1103/PhysRevLett.64.1196
  • Elaydi S. An introduction to difference equations. New York: Springer-Verlag; 1996.
  • Lynch S. Dynamical systems with applications using mathematica. Boston: Birkhäuser; 2007.
  • Luo XS, Chen G, Wang BH, et al. Hybrid control of period-doubling bifurcation and chaos in discrete nonlinear dynamical systems. Chaos Solit Fractals. 2003;18(4):775–783. doi: 10.1016/S0960-0779(03)00028-6

Appendices

Appendix 1.

Expression for the coefficients quantities of system (41)

(A1) Ω111=(1+h)(x+y)2+h2sx2yhsy2(1+hx)2(x+y)2,Ω121=hsx2(1+hx)(x+y)2,Ω131=h(x+y)3h2(x+y)3h3sx3y+hsy2+3h2sxy2+h2sy3(1+hx)3(x+y)3,Ω141=h2sx2(xy)2hsxy(1+hx)2(x+y)3,Ω151=hsx2(1+hx)(x+y)3,Ω112=hδy2(x+y)2,Ω122=1hrδ+hδx2(x+y)2,Ω132=hδy2(x+y)3,Ω142=2hδxy(x+y)3,Ω152=hδx2(x+y)3.(A1)

Appendix 2.

Expression for the coefficients quantities of system (44)

(A2) η=12(1hrs(r1)1+h+hs(r1)+1+hrδ(r1)),ζ=124((1hrs(r1)1+h+hs(r1))(1+hrδ(r1))+h2r2(r1)21+h+hs(r1))(1hrs(r1)1+h+hs(r1)+1+hrδ(r1))2,r11=Ω131Ω121+ηΩ141Ω111Ω141η2Ω151Ω121+(Ω111)2Ω151Ω1212ηΩ111Ω151Ω121,r12=ζΩ1412ζηΩ151Ω121+2ζΩ111Ω151Ω121,r13=ζ2Ω151Ω121,r21=1ζ((ηΩ111)Ω131Ω121+(ηΩ111)2Ω141(ηΩ111)3Ω151Ω121Ω132(Ω121)2Ω142Ω121(ηΩ111)Ω152(ηΩ111)2),r22=(ηΩ111)Ω1412(ηΩ111)Ω121Ω151+Ω142Ω121+2Ω152(ηΩ111),r23=ζΩ151(ηΩ111)Ω121ζΩ152.(A2)

Appendix 3.

Expression for the coefficients quantities of system (48)

(A3) Ω111=(1+h)(x+y)2+h2sx2yhsy2(1+hx)2(x+y)2,Ω121=hsx2(1+hx)(x+y)2,Ω131=h(x+y)3h2(x+y)3h3sx3y+hsy2+3h2sxy2+h2sy3(1+hx)3(x+y)3,Ω141=h2sx2(xy)2hsxy(1+hx)2(x+y)3,Ω151=hsx2(1+hx)(x+y)3,Ω112=hδy2(x+y)2,Ω122=1hrδ+hδx2(x+y)2,Ω132=hδy2(x+y)3,Ω142=2hδxy(x+y)3,Ω152=hδx2(x+y)3,k1=hy2(x+y)2,k2=hr+hx2(x+y)2,k3=hy2(x+y)3,k4=2hxy(x+y)3,k5=hx2(x+y)3.(A3)