1,950
Views
22
CrossRef citations to date
0
Altmetric
Articles

Bifurcation analysis and chaos control for a plant–herbivore model with weak predator functional response

ORCID Icon, , &
Pages 481-501 | Received 24 Feb 2019, Accepted 27 Jun 2019, Published online: 09 Jul 2019

Abstract

The interaction between plants and herbivores is one of the most fundamental processes in ecology. Discrete-time models are frequently used for describing the dynamics of plants and herbivores interaction with non-overlapping generations, such that a new generation replaces the old at regular time intervals. Keeping in view the interaction of the apple twig borer and the grape vine, the qualitative behaviour of a discrete-time plant–herbivore model is investigated with weak predator functional response. The topological classification of equilibria is investigated. It is proved that the boundary equilibrium undergoes transcritical bifurcation, whereas unique positive steady-state of discrete-time plant–herbivore model undergoes Neimark–Sacker bifurcation. Numerical simulation is provided to strengthen our theoretical discussion.

2010 AMS Mathematics Subject Classifications:

1. Introduction

In mathematical biology, plant–herbivore models are basically modifications of prey–predator models [Citation8]. The interaction between plants and herbivores has been investigated by many researchers both in differential and difference equations. Kartal [Citation31] investigated the dynamical behaviour of a plant–herbivore model including both differential and difference equations. In [Citation30] Kang et al. discussed bistability, bifurcation and chaos control in a discrete-time plant–herbivore model. In [Citation38] Liu et al. investigated stability, limit cycle, Neimark–Sacker bifurcations and homoclinic bifurcation for a plant–herbivore model with toxin-determined functional response. Li et al. [Citation37] discussed period-doubling and Neimark–Sacker bifurcations for a plant–herbivore model incorporating plant toxicity in the functional response of plant–herbivore interactions. Similarly, for some other discussions related to qualitative behaviour of plant–herbivore models, the interested reader is referred to [Citation2,Citation10,Citation21,Citation25–27,Citation29,Citation33,Citation45,Citation47,Citation52] and references are therein.

Taking into account the interaction of the apple twig borer and the grape vine [Citation3], a mathematical model for plant–herbivore interaction based on weak predator functional response is given by (1) xn+1=xnα(1+yn2)+βxn,yn+1=γyn(1+xn),(1) where xn and yn are population densities for grape vine and apple twig borer, respectively. Moreover, α, β and γ are positive parameters. For some earlier investigations related to model (Equation1), we refer to [Citation4–7,Citation34,Citation35]. Moreover, taking into account strong predator functional response, Din [Citation12] discussed global stability of system (Equation1), and Khan et al. [Citation32] investigated Neimark–Sacker bifurcation for system (Equation1) with strong predator functional response.

The main findings of this paper are summarized as follows:

  • Existence of equilibria and their stability analysis is investigated.

  • It is proved that model (Equation1) undergoes transcritical bifurcation at its boundary equilibrium by implementing bifurcation theory of normal forms and centre manifold theorem.

  • Direction and existence criteria for Neimark–Sacker bifurcation are investigated at positive steady-state of plant–herbivore model (Equation1).

  • Pole-placement and hybrid control methods are implemented in order to discuss chaos control in system (Equation1).

Moreover, in Section 2 the existence of steady-states and their local asymptotic behaviour is analysed. Section 3 is dedicated to investigate transcritical bifurcation about boundary steady-state of system (Equation1). In Section 4, we discuss Neimark–Sacker bifurcation at positive steady-state of model (Equation1). We study pole-placement chaos control and hybrid control strategies in Section 5. Finally, in Section 6, numerical simulations are carried out to authenticate the theoretical discussion.

2. Existence of equilibria and stability

In this section, first we investigate fixed points for system (Equation1). For this, these fixed points must solve the following algebraic system: (2) x=xα(1+y2)+βx,y=γ(1+x)y.(2) Simple computation yields the following steady-states for the plant–herbivore model (Equation1): P0=(0,0),P1=(1αβ,0),P=(1γγ,γ(1+β)(αγ+β)αγ). Assume that 0<α<1,β>0,β/(1α+β)<γ<1, then P is unique positive equilibrium point of the system (Equation1). For β=0.2, the region in αγ-plane where inequalities 0<α<1,β/(1α+β)<γ<1 are satisfied is depicted in Figure  by red shaded region. In order to see the dynamical behaviour of system (Equation1) at equilibria, we first compute Jacobian matrix of system (Equation1) at (x,y) as follows: (3) J(x,y):=(α(1+y2)(α(1+y2)+βx)22αxy(α(1+y2)+βx)2γyγ(1+x)).(3) Then, at (x,y)=(0,0) the Jacobian matrix J(x,y) given in (Equation3) reduces to J(0,0):=(1α00γ). Then topological classification of P0 is given as follows:

  • P0 is a sink if and only if α>1 and 0<γ<1.

  • P0 is a saddle point if and only if 0<α<1 and 0<γ<1, or α>1 and γ>1.

  • P0 is a source if and only if 0<α<1 and γ>1.

  • P0 is a non-hyperbolic point if and only if α=1, or γ=1.

Figure 1. Existence region (red) for P at β=0.2.

Figure 1. Existence region (red) for P∗ at β=0.2.

Next, for α[0,2] and γ[0,2] the aforementioned topological classification for P0 is depicted in Figure . Moreover, for the existence of boundary equilibrium P1, we assume that 0<α<1, then the Jacobian matrix J(x,y) at P1 is computed as follows: J(P1):=(α00γ(1α+β)β). Furthermore, topological classification for boundary equilibrium P1 is itemized as follows:

  • P1 is a sink if and only if 0<α<1 and 0<γ<β/(1α+β).

  • P1 is a saddle point if and only if 0<α<1 and γ>β/(1α+β).

  • P1 is a non-hyperbolic point if and only if γ=β/(1α+β).

Figure 2. Topological classification of P0 for α[0,2] and γ[0,2].

Figure 2. Topological classification of P0 for α∈[0,2] and γ∈[0,2].

Furthermore, for α[0,1], β[0,200] and γ=0.995 the topological classification of P1 is depicted in Figure . At the end of this section, we explore dynamics of model (Equation1) at its unique positive steady-state P. For this, first Jacobian matrix at this equilibrium is computed as follows: J(P):=(1+ββγ2(γ1)α(β(γ1)+γαγ)γ3/2γ(β(γ1)+γαγ)α1). Moreover, characteristic polynomial for J(P) is computed as follows: (4) F(λ)=λ2(2β(1γ1))λ+3+5β2α(1γ)3βγ2γ(1+β).(4) From (Equation4) we have F(1)=2(1γ)(β(γ1)+γαγ)γ and F(1)=6+6β2α(1γ)4βγ2(1+β)γ. Assume that 0<α<1,β>0, and β/(1α+β)<γ<1, then from first part of last inequality we have β<γ(1α+β), or equivalently β(γ1)+γαγ>0. Hence, it follows that F(1)>0. Similarly, one can prove that F(1)>0 under the existence conditions for positive steady-state P of system (Equation1). That is, F(1)>0 for 0<α<1,β>0, and β/(1α+β)<γ<1. Indeed, it follows that F(1)=6+6β2α(1γ)4βγ2(1+β)γ=2+6β2α(1γ)4βγ2(1+β)γ+4=2(1γ)(1α+β)4(1γ)(βγ)+4>2(1γ)(βγ)4(1γ)(βγ)+4=2(2+ββγ)>2(1+α)>0. Therefore, according to Jury condition roots of F(λ)=0 inside the unit open disk if and only if F(0)<1, which on simplification gives that 2α+β>2, or γ<3β/(22α+2β) and 2α+β2. Similarly, F(0)>1 if and only if 2α+β<2 and 3β/(22α+2β)<γ. Moreover, F(0)=1 if and only if γ=3β/(22α+2β) and 2α+β<2. Due to aforementioned computations, we have the following results:

Lemma 2.1

Assume that 0<α<1,β>0, and β/(1α+β)<γ<1, then the following conditions hold true:

  • P is a sink if and only if 2α+β>2, or γ<3β/(22α+2β) and 2α+β2.

  • P is a source if and only if 2α+β<2 and 3β/(22α+2β)<γ.

  • P is a non-hyperbolic fixed point if and only if γ=3β/(22α+2β) and 2α+β<2.

In Figure , topological classification of P is depicted for 0<α<1, 0<γ<1 and β=0.1.

Figure 3. Topological classification of P1 for α[0,1], β[0,200] and γ=0.995.

Figure 3. Topological classification of P1 for α∈[0,1], β∈[0,200] and γ=0.995.

Figure 4. Topological classification of P for 0<α<1, 0<γ<1 and β=0.1.

Figure 4. Topological classification of P∗ for 0<α<1, 0<γ<1 and β=0.1.

3. Transcritical bifurcation

In this section, we investigate that boundary equilibrium P1 undergoes transcritical bifurcation. For this, we assume that γγ0:=β1α+β. Consider the set ΩTB given by ΩTB:={(α,β,γ0)R+3:γ0=β1α+β,0<α<1,β>0}. Assume that (α,β,γ0)ΩTB, then system (Equation1) is equivalently represented by the following 2-dimensional map: (5) (uv)(uα(1+v2)+βu(γ0+γ¯)v(1+u)),(5) where γ¯ is very small perturbation in γ0. Furthermore, if we consider x=u(1α)/β and y=v, then map (Equation5) is transformed into the following map: (6) (xy)(x+1αβα(1+y2)+β(x+1αβ)(γ0+γ¯)y(1+x+1αβ)).(6) An application of Taylor series expansion about (x,y,γ¯)=(0,0,0) yields that (7) (xy)(α00γ0(1α+β)β)(xy)+(f(x,y)g(x,y,γ¯)),(7) where f(x,y)=αβx2α(1α)βy2+αβ2x3+α(12α)xy2+O((|x|+|y|)4) and g(x,y,γ¯)=γ0xy+(1α+ββ)yγ¯+xyγ¯. Since at γ0=β/(1α+β) the linear part of map (Equation7) is already in canonical form. Therefore, in order to implement the centre manifold theorem [Citation9], let WC(0,0,0) denotes the centre manifold for the map (Equation7) which is evaluated at (0,0) in a small neighbourhood of γ¯=0. Then, WC(0,0,0) is computed as follows: WC(0,0,0)={(x,y,γ¯)R3:y=m1x2+m2xγ¯+m3γ¯2+O((|x|+|γ¯|)3)}, where m1=m2=m3=0. Moreover, we define the following mapping which is restricted to the centre manifold WC(0,0,0): F:xx+γ¯+s1x2+s2xγ¯+s3γ¯2+O((|x|+|γ¯|)3), where s1=αβ,s2=s3=0. Furthermore, it follows that F(0,0)=0,Fx(0,0)=1,Fγ¯(0,0)=1,Fxx(0,0)=2αβ<0. Then, the following theorem gives the parametric conditions for existence and direction of transcritical bifurcation for system (Equation1) at its boundary fixed point P1.

Theorem 3.1

Suppose that γ=β/(1α+β) and 0<α<1, then system (Equation1) undergoes transcritical bifurcation at its boundary steady-state P1 when parameter γ varies in a small neighbourhood of γ0=β/(1α+β). Furthermore, two steady-states bifurcate from P1 for γ<γ0, and merge as the steady-state P1 at γ=γ0 and disappear at γ>γ0.

4. Neimark–Sacker bifurcation

In this section, we study that unique positive steady-state of system (Equation1) undergoes Neimark–Sacker bifurcation. For this, necessary conditions for existence of Neimark–Sacker bifurcation at P are given as follows: γ=3β22α+2β,2α+β<2. Next, we consider the following set: SNB:={(α,β,γ1)R+3:γ1=3β22α+2β,2α+β<2,0<α<1,β>0}. Assume that (α,β,γ1)SNB, and γ~ be small perturbation in γ1, then system (Equation1) can be expressed by the following two-dimensional perturbed map: (8) (uv)(uα(1+v2)+βu(γ1+γ~)v(1+u)).(8) In order to translate the unique positive fixed point (1(γ1+γ~)γ1+γ~,(γ1+γ~)(1+β)(α(γ1+γ~)+β)α(γ1+γ~)) of the perturbed map (Equation8) at origin, we consider the following translations: (9) x=u1(γ1+γ~)γ1+γ~,y=v(γ1+γ~)(1+β)(α(γ1+γ~)+β)α(γ1+γ~).(9) Then, from (Equation8) and (Equation9) it follows that (10) (xy)(a11a12a21a22)(xy)+(f2(x,y)g2(x,y)),(10) where (a11a12a21a22)=(1+ββγ1+γ~(γ1+γ~)(β((γ1+γ~)1)+(γ1+γ~)α(γ1+γ~))α×2((γ1+γ~)1)α(β((γ1+γ~)1)+(γ1+γ~)α(γ1+γ~))(γ1+γ~)3/21),f2(x,y)=(αβ+αβY2(α+βX+αY2)3)x2(2αY(αβX+αY2)(α+βX+αY2)3)xy(X(α2+αβX3α2Y2)(α+βX+αY2)3)y2+(αβ2+αβ2Y2(α+βX+αY2)4)x3+(2αY(2αββ2X+2αβY2)(α+βX+αY2)4)x2y+(αβ2X2α38α2βXY2+3α3Y4+2α3Y2(α+βX+αY2)4)xy2+(4X(α2βXYα3Y3+α3Y)(α+βX+αY2)4)y3+O((|x|+|y|)4),g2(x,y)=(γ1+γ~)xy,X=1(γ1+γ~)γ1+γ~,Y=(γ1+γ~)(1+β)(α(γ1+γ~)+β)α(γ1+γ~). Moreover, the characteristic polynomial for (a11a12a21a22) is computed as follows: (11) P(ρ)=ρ2(2β(1γ1+γ~1))ρ+3+5β2α(1(γ1+γ~))3βγ1+γ~2(γ1+γ~)(1+β).(11) Assume that ρ(γ~) and ρ¯(γ~) be complex conjugate roots of (Equation11), then a simple computation yields that |ρ(γ~)|=32α+2β+2(α1β)γ~6β(1α+β)3β+2(1α+β)γ~. Furthermore, it follows that |ρ(0)|=1, but ρi(0)1 for all i=1,2,3,4 if and only if (12) 13(4+2α+β)2,0,1,2.(12) Since 13(4+2α+β)<0 and (α,β,γ1)SNB, therefore 13(4+2α+β)2. Hence, condition (Equation12) is satisfied. Moreover, it follows that (d|ρ(γ~)|dγ~)γ~=0=(1α+β)(22αβ)3β>0. The following similarity transformation is considered in order to convert linear part of (Equation10) into canonical form at γ~=0: (13) (xy)=(a120ϑa11φ)(wz),(13) where ϑ:=4+2α+β6,φ:=(22αβ)(10+2α+β)6 and a11=13(1+2α+β),a12=2α1α+βα(2α2+β)33β. Then, from transformation (Equation13) it follows that (14) (wz)=(1a120ϑa11φa121φ)(xy).(14) Assuming γ~=0, then from (Equation10), (Equation13) and (Equation14) we obtain the following normal form of map (Equation10): (15) (wz)=(ϑφφϑ)(wz)+(F(w,z)G(w,z)),(15) where F(w,z):=(33ββ1α+β2αβ(22αβ))f2(a12w,(ϑa11)wφz) and G(w,z):=3(4αβg2(a12w,(ϑa11)wφz)+3ββ1α+βf2(a12w,(ϑa11)wφz))2αβ(22αβ)(10+2α+β). Taking into account the bifurcation theory of normal forms [Citation28,Citation36,Citation43,Citation48,Citation49], the first Lyapunov exponent at (w,z)=(0,0) is computed as follows: L=Re((12ρ(0))ρ¯(0)21ρ(0)τ20τ11)12|τ11|2|τ02|2+Re(ρ¯(0)τ21), where τ20=18[FwwFzz+2Gwz+i(GwwGzz2Fwz)],τ11=14[Fww+Fzz+i(Gww+Gzz)],τ02=18[FwwFzz2Gwz+i(GwwGzz+2Fwz)], and τ21=116[Fwww+Fwzz+Gwwz+Gzzz+i(Gwww+GwzzFwwzFzzz)]. Due to aforementioned computations, we have the following theorem:

Theorem 4.1

Suppose that L0, 0<α<1, β/(1α+β)<γ<1 and 2α+β<2, then unique positive equilibrium point P of system (Equation1) undergoes Neimark–Sacker bifurcation when the bifurcation parameter γ varies in a small neighbourhood of γ1=3β/(22α+2β). Moreover, if L<0, then an attracting invariant closed curve bifurcates from the equilibrium point for γ>γ1, and if L>0, then a repelling invariant closed curve bifurcates from the equilibrium point for γ<γ1.

5. Chaos control

Chaos control is a method of stabilization with the help of small perturbations which are applied to unstable periodic orbits for a given system. The purpose of chaos control is to make chaotic behaviour more predictable and stable. For this, a small perturbation is applied as compare to the original size of the system under consideration, and in a result the natural dynamics of the system is prevented from any major modification. Recently, chaos control for irregular complex dynamics has developed as one of the leading topics in nonlinear applied science. The pioneer articles related to chaos control were published in 1990 and after that their number has been steadily increasing [Citation46]. Moreover, the practical methods related to chaos control can be implemented in various areas such as communications, physics laboratories, biochemistry, turbulence, and cardiology [Citation40]. Furthermore, in the case of dynamical systems related to biological breeding of species, chaos control is considered to be an important feature for investigation. On the other hand, population models with non-overlapping generations have more irregular complex behaviour. For some recent investigation related to chaos control in discrete-time models we refer to [Citation1,Citation13–20,Citation22–24] and references are therein.

In this section, first we discuss pole-placement chaos control method based on state feedback control which was introduced by Romeiras et al. [Citation44] (also see [Citation41]). This method is also known as generalized or modified OGY method proposed by Ott et al. [Citation42]. For the application of pole-placement technique to model (Equation1), one can rewrite this system as follows: (16) xn+1=xnα(1+yn2)+βxn=f(xn,yn,γ),yn+1=γyn(1+xn)=g(xn,yn,γ),(16) where γ denotes parameter for chaos control. Furthermore, it is assumed that γ lies in a small interval of type |γγ0|<δ, where δ>0 and γ0 represents the nominal parameter lies in the chaotic region. One can apply pole-placement method in order to move unstable trajectory towards desired stable orbit. For this, suppose that (x,y) be an unstable fixed point for model (Equation1) lies in a chaotic region under the influence of transcritical or Neimark–Sacker bifurcations. Keeping in view the method of linearization, one can approximate system (Equation16) in the neighbourhood of the unstable fixed point (x,y) as follows: (17) [xn+1xyn+1y]A[xnxyny]+B[γγ0],(17) where A=[f(x,y,γ0)xnf(x,y,γ0)yng(x,y,γ0)yng(x,y,γ0)yn],B=[f(x,y,γ0)γg(x,y,γ0)γ]=[0y(1+x)]. In order to check that system (Equation16) is controllable, the following matrix is computed: (18) C=[B:AB]=[0(f(x,y,γ0)yn)(y(1+x))y(1+x)(g(x,y,γ0)yn)(y(1+x))].(18) Assume that (x,y) is positive fixed point, then it follows that y(1+x)0. Therefore, rank of C is 2 at positive equilibrium point. Next, we assume that [γγ0]=K[xnxyny], where K=[k1k2], then system (Equation17) can be written as (19) [xn+1xyn+1y][ABK][xnxyny].(19) Moreover, equilibrium point (x,y) is a sink if and only if both eigenvalues μ1 and μ2 for the matrix ABK lie in an open unit disk. Moreover, μ1 and μ2 are called regulator poles. Placing these eigenvalues at appropriate position is called pole-placement method. Furthermore, pole-placement problem has unique solution because rank of matrix C is two. Denote λ2+α1λ+α2=0 as characteristic equation for matrix A and μ2+β1μ+β2=0 is characteristic equation of ABK, then unique solution of the pole-placement problem is given as follows: (20) K=[β2α2β1α1]T1,(20) where T=CM and M=[α1110]. Then it follows that (21) T=CM=[fyngγ0α1gγ+gyngγgγ],(21) where all partial derivatives in (Equation21) are evaluated at (x,y,γ0). From (Equation20) and (Equation21), we have the following unique solution of pole-placement problem: k1=β2α2(f(x,y,γ0)yn)(g(x,y,γ0)c)(β1α1)(α1+g(x,y,γ0)yn)(f(x,y,γ0)yn)(g(x,y,γ0)γ),k2=β1α1g(x,y,γ0)γ. Secondly, we apply another comparatively simple chaos control method known as hybrid control strategy based on parameter perturbation and state feedback control [Citation39,Citation50]. An application of hybrid control strategy to model (Equation1) yields the following controlled system: (22) xn+1=κ(xnα(1+yn2)+βxn)+(1κ)xn,yn+1=κ(γyn(1+xn))+(1κ)yn,(22) where 0<κ<1 is control parameter. Assume that (x,y) be an equilibrium point of system (Equation22), then Jacobian matrix for system (Equation22) computed at (x,y) is given as follows: (23) J(x,y):=(κα(1+(y)2)(α(1+(y)2)+βx)2+1κ2καxy(α(1+(y)2)+βx)2γκyγκ(1+x)+1κ).(23) It is easy to observe that system (Equation22) is controllable as long as the steady-state (x,y) of system (Equation22) is locally asymptotically stable. Keeping in view this fact, one has the following result for controllability of system (Equation22).

Lemma 5.1

The equilibrium point (x,y) of system (Equation22) is locally asymptotically stable if the following holds true: |TrJ(x,y)|<1+detJ(x,y)<2.

6. Numerical simulation and discussion

First, we choose α=0.99, β=0.008, γ[0.2,0.7] and (x0,y0)=(1.25,0.001). Then, system (Equation1) undergoes transcritical bifurcation at its boundary equilibrium point (1.25,0) as bifurcation parameter γ passes through γ0.4444. Furthermore, at γ0.666667 the positive steady-state (0.499999,0.0778499) exists and undergoes Neimark–Sacker bifurcation. At (α,β,γ)=(0.99,0.008,0.666667) the characteristic equation for Jacobian matrix of model (Equation1) is computed as follows: λ21.996λ+1=0. Obviously, λ1=0.9980.0632139i and λ2=0.998+0.0632139i be roots for aforementioned characteristic equation with |λ1,2|=1. The maximum Lyapunov exponents (MLE) and bifurcation diagrams are depicted in Figure . Secondly, we select α=0.2, β=0.5, γ[0.45,0.95] and (x0,y0)=(0.78,1.426), then model (Equation1) undergoes Neimark–Sacker bifurcation at its positive steady-state (0.733333,1.47196) as bifurcation parameter γ varies in a small neighbourhood of γ=0.576923. For the parametric values (α,β,γ)=(0.2,0.5,0.576923) the characteristic equation of the Jacobian matrix of system (Equation1) is computed as follows: λ21.63333λ+1=0. Moreover, we have λ1=0.8166670.57711i and λ2=0.816667+0.57711i as roots for aforementioned characteristic equation with |λ1,2|=1. MLE and bifurcation diagrams are depicted in Figure . On the other hand, some phase portraits for γ=0.573,0.576923,0.58,0.65 are depicted in Figure .

Finally, we take α=0.2, β=0.5 and γ=0.95 to verify the effectiveness of chaos control strategies introduced in Section 5. Under this selection of parameters, system (Equation1) has unique positive steady-state given by (0.0526316,1.96683). An application of pole-placement method gives the following controlled system: (24) xn+1=xn0.2(1+yn2)+0.5xn,yn+1=(0.95k1(xn0.0526316)k2(yn1.96683))yn(1+xn).(24) Then, variational matrix for system (Equation24) is computed as follows: [0.9736840.0414071.868492.07035k112.07035k2]. Furthermore, the lines for marginal stability are computed as follows: L1:k2=0.9849550.0209795k1,L2:k2=1.42005+1.57346k1, and L3:k2=0.02532540.0425261k1. The triangular stability region bounded by these marginal stability lines is depicted in Figure .

Figure 5. Bifurcation diagrams and MLE for system (Equation1) with α=0.99, β=0.008, γ[0.2,0.7] and (x0,y0)=(1.25,0.001): (a) bifurcation diagram for xn, (b) bifurcation diagram for yn and (c) MLE.

Figure 5. Bifurcation diagrams and MLE for system (Equation1(1) xn+1=xnα(1+yn2)+βxn,yn+1=γyn(1+xn),(1) ) with α=0.99, β=0.008, γ∈[0.2,0.7] and (x0,y0)=(1.25,0.001): (a) bifurcation diagram for xn, (b) bifurcation diagram for yn and (c) MLE.

Figure 6. Bifurcation diagrams and MLE for system (Equation1) with α=0.2, β=0.5, γ[0.45,0.95] and (x0,y0)=(0.78,1.426): (a) bifurcation diagram for xn, (b) bifurcation diagram for yn and (c) MLE.

Figure 6. Bifurcation diagrams and MLE for system (Equation1(1) xn+1=xnα(1+yn2)+βxn,yn+1=γyn(1+xn),(1) ) with α=0.2, β=0.5, γ∈[0.45,0.95] and (x0,y0)=(0.78,1.426): (a) bifurcation diagram for xn, (b) bifurcation diagram for yn and (c) MLE.

Figure 7. Phase portraits of system (Equation1) for α=0.2, β=0.5, x0=0.78,y0=1.426 and with different values of γ: (a) phase portrait for γ=0.573, (b) phase portrait for γ=0.576923, (c) phase portrait for γ=0.58 and (d) phase portrait for γ=0.65.

Figure 7. Phase portraits of system (Equation1(1) xn+1=xnα(1+yn2)+βxn,yn+1=γyn(1+xn),(1) ) for α=0.2, β=0.5, x0=0.78,y0=1.426 and with different values of γ: (a) phase portrait for γ=0.573, (b) phase portrait for γ=0.576923, (c) phase portrait for γ=0.58 and (d) phase portrait for γ=0.65.

Figure 8. Bounded stability region for system (Equation24).

Figure 8. Bounded stability region for system (Equation24(24) xn+1=xn0.2(1+yn2)+0.5xn,yn+1=(0.95−k1(xn−0.0526316)−k2(yn−1.96683))yn(1+xn).(24) ).

Next, for same parametric values we apply hybrid control method to system (Equation1). In this case system (Equation22) is written as follows: (25) xn+1=κ(xn0.2(1+yn2)+0.5xn)+(1κ)xn,yn+1=κ(0.95yn(1+xn))+(1κ)yn.(25) Then, Jacobian matrix of system (Equation25) and its characteristic equation are computed as follows: [10.0263158κ0.041407κ1.86849κ1] and λ2(20.0263158κ)λ+10.0263158κ+0.0773684κ2=0. Keeping in view Jury condition and aforementioned characteristic equation, |20.0263158κ|<20.0263158κ+0.0773684κ2<2 if and only if 0<κ<0.340136. At κ=0.33 the plots for system (Equation25) are depicted in Figure .

Figure 9. Plots for system (Equation25) with κ=0.33 and (x0,y0)=(0.0526316,1.96683): (a) plot for xn, (b) plot for yn and (c) phase portrait.

Figure 9. Plots for system (Equation25(25) xn+1=κ(xn0.2(1+yn2)+0.5xn)+(1−κ)xn,yn+1=κ(0.95yn(1+xn))+(1−κ)yn.(25) ) with κ=0.33 and (x0,y0)=(0.0526316,1.96683): (a) plot for xn, (b) plot for yn and (c) phase portrait.

7. Concluding remarks

Local asymptotic stability, bifurcation analysis and chaos control are investigated for the apple twig borer and the grape vine type plant–herbivore model. The discrete-time model is obtained by taking into account the weak predator functional response. Furthermore, existence of equilibria and their stability analysis is investigated. Taking into account the bifurcating behaviour for biological populations, it must be noted that these phenomena are essential for competition between plants and herbivores [Citation11]. It is proved that plant–herbivore model undergoes transcritical bifurcation at its boundary equilibrium by implementing bifurcation theory of normal forms and centre manifold theorem. Direction and existence criteria for Neimark–Sacker bifurcation are investigated at positive steady-state of plant–herbivore model. Bifurcating behaviour and chaos have always been considered as disadvantageous phenomena in biology. On the other hand, due to chaotic behaviour population can undergo a higher risk of extinction due to the unpredictability. Therefore, these are catastrophic for the breeding of biological population [Citation51]. In order to prevent biological populations from this ruinous situation, one might think about applications of chaos and bifurcation control. Pole-placement and hybrid control methods are implemented in order to discuss chaos control in the system. Furthermore, our investigation reveals that the pole-placement method is more effective as compared to the hybrid control method.

8. Future work

It must be noted that the weak predator functional response  in system (1) is similar to Holling type-III function Equation1. Therefore, it is more interesting to apply Holling type-II functional response. With such implementation of Holling type-II functional response, plant–herbivore model (Equation1) can be rewritten as follows: (26) xn+1=xnα(1+yn)+βxn,yn+1=γyn(1+xn).(26) Stability, bifurcation analysis and chaos control for model (Equation26) is our future work for investigation.

Acknowledgements

The authors are grateful to the referees for their excellent suggestions which greatly improve the presentation of the paper.

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • M.A. Abbasi and Q. Din, Under the influence of crowding effects: Stability, bifurcation and chaos control for a discrete-time predator–prey model, Int. J. Biomath. 12(4) (2019), p. 1950044.
  • K.C. Abbott and G. Dwyer, Food limitation and insect outbreaks: Complex dynamics in plant–herbivore models, J. Anim. Ecol. 76 (2007), pp. 1004–1014.
  • L.J.S. Allen, M.K. Hannigan, and M.J. Strauss, Mathematical analysis of a model for a plant–herbivore system, Bull. Math. Biol. 55(4) (1993), pp. 847–864.
  • L.J.S. Allen, M.J. Strauss, H.G. Thorvilson, and W.N. Lipe, A preliminary mathematical model of the apple twig borer (Coleoptera: Bostrichidae) and grapes on the Texas High Plains, Ecol. Model. 58 (1991), pp. 369–382.
  • R. Beiriger, The bionomics of Amphicerus bicaudatus (say), the apple twig borer, on the Texas High Plains, M.Sc. thesis, Texas Tech University, Lubbock, TX, 1988.
  • R. Beiriger, H. Thorvilson, W. Eipe, and D. Rummel, The life history of the apple twig borer, Proceedings of the Texas Grape Growers Association, Vol. 12, 1988, pp. 101–106.
  • R. Beiriger, H. Thorvilson, W. Lipe, and D. Rummel, Determination of adult apple twig borer sex based on a comparison of internal and external morphology, Southwest. Entomol. 14 (1989), pp. 199–203.
  • Y.M. Buckley, M. Rees, A.W. Sheppard, and M.J. Smyth, Stable coexistence of an invasive plant and biocontrol agent: A parameterized coupled plant–herbivore model, J. Appl. Ecol. 42 (2005), pp. 70–79.
  • J. Carr, Application of Center Manifold Theory, Springer-Verlag, New York, 1981.
  • C. Castillo-Chavez, Z. Feng, and W. Huang, Global dynamics of a plant–herbivore model with toxin-determined functional response, SIAM J. Appl. Math. 72(4) (2012), pp. 1002–1020.
  • Q. Cui, Q. Zhang, Z. Qiu, and Z. Hu, Complex dynamics of a discrete-time predator–prey system with Holling IV functional response, Chaos Soliton Fract. 87 (2016), pp. 158–171.
  • Q. Din, Global behavior of a plant–herbivore model, Adv. Differ. Equ. 2015 (2015), p. 119.
  • Q. Din, Complexity and chaos control in a discrete-time prey–predator model, Commun. Nonlinear Sci. Numer. Simul. 49 (2017), pp. 113–134.
  • Q. Din, Neimark–Sacker bifurcation and chaos control in Hassell–Varley model, J. Differ. Equ. Appl. 23(4) (2017), pp. 741–762.
  • Q. Din, Bifurcation analysis and chaos control in discrete-time glycolysis models, J. Math. Chem. 56(3) (2018), pp. 904–931.
  • Q. Din, Controlling chaos in a discrete-time prey–predator model with Allee effects, Int. J. Dynam. Control 6(2) (2018), pp. 858–872.
  • Q. Din, Qualitative analysis and chaos control in a density-dependent host–parasitoid system, Int. J. Dynam. Control 6(2) (2018), pp. 778–798.
  • Q. Din, A novel chaos control strategy for discrete-time Brusselator models, J. Math. Chem. 56(10) (2018), pp. 3045–3075.
  • Q. Din, Stability, bifurcation analysis and chaos control for a predator–prey system, J. Vib. Control 25(3) (2019), pp. 612–626.
  • Q. Din, T. Donchev, and D. Kolev, Stability: Bifurcation analysis and chaos control in chlorine dioxide–iodine–malonic acid reaction, MATCH Commun. Math. Comput. Chem. 79(3) (2018), pp. 577–606.
  • Q. Din, A.A. Elsadany, and H. Khalil, Neimark–Sacker bifurcation and chaos control in a fractional-order plant–herbivore model, Discrete Dyn. Nat. Soc. 2017 (2017), pp. 1–15.
  • Q. Din, Ö.A. Gümüş, and H. Khalil, Neimark–Sacker bifurcation and chaotic behaviour of a modified host–parasitoid model, Z. Naturforsch. A 72(1) (2017), pp. 25–37.
  • Q. Din and M. Hussain, Controlling chaos and Neimark–Sacker bifurcation in a host–parasitoid model, Asian J. Control 21(3) (2019), pp. 1202–1215.
  • Q. Din and M.A. Iqbal, Bifurcation analysis and chaos control for a discrete-time enzyme model, Z. Naturforsch. A 74(1) (2019), pp. 1–14.
  • M. El-Shahed, A.M. Ahmed, and I.M.E. Abdelstar, Dynamics of a plant–herbivore model with fractional order, Progr. Fract. Differ. Appl. 3(1) (2017), pp. 59–67.
  • Z. Feng and D.L. DeAngelis, Mathematical Models of Plant–Herbivore Interactions, CRC Press, Taylor and Francis Group, Boca Raton, 2018.
  • Z. Feng, Z. Qiu, R. Liu, and D.L. DeAngelis, Dynamics of a plant–herbivore–predator system with plant-toxicity, Math. Biosci. 229(2) (2011), pp. 190–204.
  • J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, Springer-Verlag, New York, 1983.
  • S.S. Jothi and M. Gunasekaran, Chaos and bifurcation analysis of plant–herbivore system with intra-specific competitions, Int. J. Adv. Res. 3(8) (2015), pp. 1359–1363.
  • Y. Kang, D. Armbruster, and Y. Kuang, Dynamics of a plant–herbivore model, J. Biol. Dyn. 2(2) (2008), pp. 89–101.
  • S. Kartal, Dynamics of a plant–herbivore model with differential–difference equations, Cogent Math. 3(1) (2016), p. 1136198.
  • A.Q. Khan, J. Ma, and D. Xiao, Bifurcations of a two-dimensional discrete time plant–herbivore system, Commun. Nonlinear Sci. Numer. Simul. 39 (2016), pp. 185–198.
  • Y. Kuang, J. Huisman, and J.J. Elser, Stoichiometric plant–herbivore models and their interpretation, Math. Biosci. Eng. 1(2) (2004), pp. 215–222.
  • M.R.S. Kulenović and G. Ladas, Dynamics of Second Order Rational Difference Equations: With Open Problems and Conjectures, Chapman and Hall/CRC, New York, 2002.
  • M.R.S. Kulenović and O. Merino, Discrete Dynamical Systems and Difference Equations with Mathematica, Chapman and Hall/CRC, New York, 2002.
  • Y.A. Kuznetsov, Elements of Applied Bifurcation Theory, Springer-Verlag, New York, 1997.
  • Y. Li, Z. Feng, R. Swihart, J. Bryant, and H. Huntley, Modeling plant toxicity on plant–herbivore dynamics, J. Dynam. Differ. Equ. 18(4) (2006), pp. 1021–1024.
  • R. Liu, Z. Feng, H. Zhu, and D.L. DeAngelis, Bifurcation analysis of a plant–herbivore model with toxin-determined functional response, J. Differ. Equ. 245 (2008), pp. 442–467.
  • X.S. Luo, G.R. Chen, B.H. Wang, and J.Q. Fang, Hybrid control of period-doubling bifurcation and chaos in discrete nonlinear dynamical systems, Chaos Soliton Fract. 18(4) (2003), pp. 775–783.
  • S. Lynch, Dynamical Systems with Applications Using Mathematica, Birkhäuser, Boston, 2007.
  • K. Ogata, Modern Control Engineering, 2nd ed., Prentice-Hall, Englewood, NJ, 1997.
  • E. Ott, C. Grebogi, and J.A. Yorke, Controlling chaos, Phys. Rev. Lett. 64(11) (1990), pp. 1196–1199.
  • C. Robinson, Dynamical Systems: Stability, Symbolic Dynamics and Chaos, Boca Raton, New York, 1999.
  • F.J. Romeiras, C. Grebogi, E. Ott, and W.P. Dayawansa, Controlling chaotic dynamical systems, Physica D 58 (1992), pp. 165–192.
  • T. Saha and M. Bandyopadhyay, Dynamical analysis of a plant–herbivore model bifurcation and global stability, J. Appl. Math. Comput. 19(1–2) (2005), pp. 327–344.
  • E. Schöll and H.G. Schuster, Handbook of Chaos Control, Wiley-VCH, Weinheim, 2007.
  • G. Sui, M. Fan, I. Loladze, and Y. Kuang, The dynamics of a stoichiometric plant–herbivore model and its discrete analog, Math. Biosci. Eng. 4(1) (2007), pp. 29–46.
  • Y.H. Wan, Computation of the stability condition for the Hopf bifurcation of diffeomorphism on R2, SIAM J. Appl. Math. 34 (1978), pp. 167–175.
  • S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Springer-Verlag, New York, 2003.
  • L-G. Yuan and Q-G. Yang, Bifurcation, invariant curve and hybrid control in a discrete-time predator–prey system, Appl. Math. Model. 39(8) (2015), pp. 2345–2362.
  • X. Zhang, Q.L. Zhang, and V. Sreeram, Bifurcation analysis and control of a discrete harvested prey–predator system with Beddington–DeAngelis functional response, J. Franklin Inst. 347 (2010), pp. 1076–1096.
  • Y. Zhao, Z. Feng, Y. Zheng, and X. Cen, Existence of limit cycles and homoclinic bifurcation in a plant–herbivore model with toxin-determined functional response, J. Differ. Equ. 258(8) (2015), pp. 2847–2872.