
Mathematical modeling and analysis of a crop-pest interacting system helps us to understand the dynamical properties of the system such as stability, bifurcations and chaos. In this article, a predator-prey type mathematical model for pest control using bio-pesticides has been analysed to study the global stability property of the interior equilibrium point. Moreover, the occurrence and orbital stability of Hopf bifurcating limit cycle solutions have been studied using ref30's conditions. Analytical and numerical results show that the interior equilibrium of the pest control model is globally asymptotically stable. Also, Hopf bifurcating occurs when the bifurcation parameter crosses the critical value, and the bifurcating periodic solution is found to be stable.
Citation: Aeshah A. Raezah, Jahangir Chowdhury, Fahad Al Basir. Global stability of the interior equilibrium and the stability of Hopf bifurcating limit cycle in a model for crop pest control[J]. AIMS Mathematics, 2024, 9(9): 24229-24246. doi: 10.3934/math.20241179
[1] | Yudan Ma, Ming Zhao, Yunfei Du . Impact of the strong Allee effect in a predator-prey model. AIMS Mathematics, 2022, 7(9): 16296-16314. doi: 10.3934/math.2022890 |
[2] | Chaoxiong Du, Wentao Huang . Hopf bifurcation problems near double positive equilibrium points for a class of quartic Kolmogorov model. AIMS Mathematics, 2023, 8(11): 26715-26730. doi: 10.3934/math.20231367 |
[3] | Yougang Wang, Anwar Zeb, Ranjit Kumar Upadhyay, A Pratap . A delayed synthetic drug transmission model with two stages of addiction and Holling Type-II functional response. AIMS Mathematics, 2021, 6(1): 1-22. doi: 10.3934/math.2021001 |
[4] | Jawdat Alebraheem . Asymptotic stability of deterministic and stochastic prey-predator models with prey herd immigration. AIMS Mathematics, 2025, 10(3): 4620-4640. doi: 10.3934/math.2025214 |
[5] | Komal Bansal, Trilok Mathur, Narinderjit Singh Sawaran Singh, Shivi Agarwal . Analysis of illegal drug transmission model using fractional delay differential equations. AIMS Mathematics, 2022, 7(10): 18173-18193. doi: 10.3934/math.20221000 |
[6] | Binfeng Xie, Na Zhang . Influence of fear effect on a Holling type III prey-predator system with the prey refuge. AIMS Mathematics, 2022, 7(2): 1811-1830. doi: 10.3934/math.2022104 |
[7] | Sahabuddin Sarwardi, Hasanur Mollah, Aeshah A. Raezah, Fahad Al Basir . Direction and stability of Hopf bifurcation in an eco-epidemic model with disease in prey and predator gestation delay using Crowley-Martin functional response. AIMS Mathematics, 2024, 9(10): 27930-27954. doi: 10.3934/math.20241356 |
[8] | Ruizhi Yang, Dan Jin, Wenlong Wang . A diffusive predator-prey model with generalist predator and time delay. AIMS Mathematics, 2022, 7(3): 4574-4591. doi: 10.3934/math.2022255 |
[9] | Yingyan Zhao, Changjin Xu, Yiya Xu, Jinting Lin, Yicheng Pang, Zixin Liu, Jianwei Shen . Mathematical exploration on control of bifurcation for a 3D predator-prey model with delay. AIMS Mathematics, 2024, 9(11): 29883-29915. doi: 10.3934/math.20241445 |
[10] | Ming Wu, Hongxing Yao . Stability and bifurcation of a delayed diffusive predator-prey model affected by toxins. AIMS Mathematics, 2023, 8(9): 21943-21967. doi: 10.3934/math.20231119 |
Mathematical modeling and analysis of a crop-pest interacting system helps us to understand the dynamical properties of the system such as stability, bifurcations and chaos. In this article, a predator-prey type mathematical model for pest control using bio-pesticides has been analysed to study the global stability property of the interior equilibrium point. Moreover, the occurrence and orbital stability of Hopf bifurcating limit cycle solutions have been studied using ref30's conditions. Analytical and numerical results show that the interior equilibrium of the pest control model is globally asymptotically stable. Also, Hopf bifurcating occurs when the bifurcation parameter crosses the critical value, and the bifurcating periodic solution is found to be stable.
The ruin of crops by pest invasions is a serious worldwide crisis not only in farming areas but also in woodland ecosystems. This issue connected with pests has been recognized since the cultivation of crops started. Approximately 42% of the world's food supply is destroyed because of pests [1]. Recently, the biological control of pests has been earning more attention among experimenters, and its practical application is also rising in the crop field. This method seeks to reduce the reliance on pesticides by emphasizing the contribution of biological control agents where living organisms are only used to control pests. Managing pests via chemical pesticides is less expensive and destroys pests rapidly but generates high environmental loss. On the other hand, natural control is a long and expensive method to use, but with very little ecological loss [2,3].
In this study, we focus on the application of biopesticides for crop pest control. A mathematical model on this system can help us to identify the main parameters. The stability of model system shows complex dynamics such as bifurcation and chaos. The global stability of equilibrium point and the stability of a limit cycle around it are important factors in understanding the behaviors that can be seen in a dynamical system. For example, in a biological system, the stability of a limit cycle can determine whether a population of organisms will grow or decline over time. In a biological system, the stability of a limit cycle can determine whether the system will oscillate periodically or exhibit unstable behavior [4,5].
In a dynamical system, a limit cycle refers to a periodic orbit that a system can exhibit [6]. The stability of a limit cycle determines how the system behaves near the limit cycle [7]. There are two types of stability for limit cycles: asymptotic stability and structural stability. Asymptotic stability means that the system tends toward the limit cycle as time goes to infinity [8]. This means that any initial conditions near the limit cycle will approach it over time. Asymptotic stability can be further classified into stable limit cycles and unstable limit cycles. A stable limit cycle is one where the system is attracted toward the limit cycle, while an unstable limit cycle is one where the system is repelled away from the limit cycle [9]. Structural stability, on the other hand, means that the qualitative behavior of the system is preserved under small perturbations to the system's parameters. In other words, if a system exhibits a limit cycle under certain conditions, then it will continue to do so under small changes to those conditions [10,11].
Mathematical modeling and analysis helps in understanding the dynamics of the system under consideration [12]. Results obtained from the modeling approach can help in proposing proper pest control strategies [13]. Dynamical properties, such as Hopf bifurcation, chaos, limit cycle etc., have seen studied by researchers [6]. Mathematical models for pest control using biopesticides have been proposed and studied by researchers [14,15,16,17]. Pest control models are developed to analyze specific dynamics, such as Hopf bifurcation via the occurrence of limit cycles [18,19,20,21]. The authors of [22] and [23] have studied the occurrences of Hopf bifurcating periodic solutions. Little research are focused on the effect of environmental fluctuations on a sterile insect release method [24]. Mathematical models for pest management using biological control strategies have been designed and analyzed by many researchers [16,25,26]. The local stability of equilibrium points and the occurrence of Hopf bifurcation has been analyzed in the articles. But there are a few articles that focus on the global stability of equilibrium points and the stability of bifurcating periodic solutions which are important properties of a dynamical system [27].
In [28], Chowdhury et al. have proposed a model for crop pest control using an integrated approach (i.e., both biopesticides and chemical pesticides are used) in an optimal manner. They have analyzed the local stability of different equilibria and provided the existence of Hopf bifurcation. But the global stability and stability of a limit cycle were not addressed. Thus, in this research, we have derived a model for pest control from the mathematical model proposed in [28] to study the dynamics of the crop-pest system in the presence of biopesticides. The model system is analyzed both analytically as well as numerically for the occurrence and stability of a Hopf bifurcating limit cycle using Poore's condition. Also, the global stability of the interior equilibrium point is analyzed using a suitable Lyapunov function.
The paper is organized as follows. In section 2, the formulation of the model is given. In section 4, we will analyze the global stability of the endemic equilibrium point and the occurence of Hopf bifurcation. The stability of the bifurcating limit cycle around coexisting equilibrium is analysed in section 5. Simulations are carried out in section 6 to substantiate the analytical findings. The final section contains a discussion of the main results to conclude the the paper.
In this research, we have considered the mathematical model established in [28] and derive the mathematical model to study the global stability of the interior equilibrium and stability of the limit cycle. We provide the derivation of the model as follows.
The following hypotheses are made to derive the mathematical model:
- As system variables, the biomass of crop is denoted as X(t), the susceptible pest population as Y(t), the infected pest as I(t) and the biopesticides (viruses) as V(t).
- Logistic growth for the biomass of the crop is considered with net growth rate r1 and carrying capacity K. α is the consumption rate of the crop biomass by susceptible pests.
- In biological control of pest, biopests (generally viruses) are sprayed on the plantation. In this process, susceptible pests are affected and become the infected ones. κ is the rate of replication of the virus by lysis and μV is the mortality rate of the biopesticide (virus). πv is the constant rate of spray of virus in the environment. λ1 is the reduction rate constant of the free virus.
- Pest consumes the crop resource at a rate α which is converted to the susceptible pest population with a maximum growth rate r(X). Here, r(X) is dependent on the density of the crop biomass and is also assumed to be a scalar multiple of the biomass of crop X due to the consumption of the crop resources limited at the maximum crop cultivation capacity [r(X) = αr2X]. αr2 is the maximum growth rate of the susceptible pest. Hence the susceptible are growing at the rate governed by the consumption of crop resources, which is growing maximally at a rate r1.
- The carrying capacity of the susceptible pest is assumed as ks. Here, ks is dependent on the carrying capacity of the crop biomass K factored with a constant term c>1, i.e., ks = cK.
Based on the above assumptions, the following model is obtained:
dXdt=r1X(1−XK)−αXS,dSdt=αr2XS(1−S+IcK)−λSV,dIdt=λSV−ξI,dVdt=πv+κξI−μvV−λ1SV, | (2.1) |
with initial conditions as
X(0)≥0,S(0)≥0,I(0)≥0,V(0)≥0. | (2.2) |
The region given below is positively invariant region for the system (2.1):
Ω={(X,S,I,V)∈R4+:0≤j≤M1,0≤S+I≤M2,0≤V≤M3}, | (2.3) |
where M1 = max{K,J(0)}, M2=αr2M1cK4ξ, and M3=πv+κξM2μv.
In this section, we analyse the existence of possible equilibrium points of the model system (2.1). Then we study the local stability properties.
Analyzing the system (2.1), we get three feasible steady states, namely,
(ⅰ) the plant-pest free equilibrium point E1(0,0,0,πvμv),
(ⅱ) the pest-free equilibrium point E2(K,0,0,πvμv), and
(ⅲ) the interior equilibrium point E∗(X∗,S∗,I∗,V∗), where,
X∗=K(r1−αS∗)r1,I∗=λS∗πvξS∗(λ1−κλ)+ξμv,V∗=πvS∗(λ1−κλ)+μv, |
and S∗ is the positive root of the following cubic equation:
Φ(S)=m1S3+m2S2+m3S+m4=0 | (3.1) |
with,
m1=α2r2ξ(κλ−λ1)m2=cα2r2Kξ(λ1−κλ)+αr2ξ{r1(λ1−κλ)−αμv}−α2r2λπvm3=cKαr2ξ{αμv−r1(λ1−κλ)}+ξαr2r1μv+αr2r1λπvm4=cr1ξ(λπv−αr2Kμv). | (3.2) |
Equation (3.1) is a cubic polynomial equation, thus a real root always exists. Also, if λ1>κλ holds then m1<0,m2>0, and m3>0. Now m4>0 holds when λπv>αr2Kμv. Hence, using Descartes' rule of signs, we have the following proposition.
Proposition 1. For λ1>κλ and λπv>αr2Kμv, there always exists a unique interior equilibrium point E∗.
Here, we check the local stability of the equilibria of the system (2.1). For this analysis, we need the Jacobian matrix of the system at any equilibrium point E(X,S,I,V) is given by
JE=[Jij]4×4=[r(1−2JK)−αS−αX00αr2S(1−S+IcK)J22−αr2XScK−λS0λV−ξ−d2u1λS0−λ1Vκξ−μv−λ1S], | (3.3) |
where J22=αr2X(1−2S+IcK)−λV.
At E1, the Jacobian matrix gives the following characteristic equation in ρ:
(ρ−r)⋅(ρ+πvλμv)⋅(ρ+ξ)(ρ+μv)=0, | (3.4) |
whose eigenvalues are r>0, −πvλμv<0,−ξ<0, and −μv<0. Thus, one eigenvalue is always positive and, consequently, the axial equilibrium, E0, is unstable.
The Jacobian matrix of the system at pest-free equilibrium point E2(K,0,0,πvμv) which satisfies the following equation:
(ρ+r)⋅(ρ−αr2K+πvλμv)⋅(ρ+ξ)(x+μv)=0. | (3.5) |
Eigenvalues of the above matrix are given as −r, αr2K−πvλμv, −ξ, and −μv. Clearly, three eigenvalues are negative and remaining eigenvalue will be negative if
αr2μvK<πvλμv. | (3.6) |
The characteristic equation of JE∗=[Jij]4×4 is
ρ4+σ1ρ3+σ2ρ2+σ3ρ+σ4=0, | (3.7) |
where
σ1=−(J11+J22+J33+J44)σ2=−J12J21+J11J22−J23J32+(J11+J22)J33−J24J42−J34J43+(J11+J22+J33)J44σ3=J11J23J32+J33(J12J21−J11J22)+J24(J11J42+J33J42)−J23J34J42−J24J32J43+J34J43(J11+J22)+J44(J12J21−J11J22+J23J32)−(J11+J22)J33J44σ4=J11J22J33J44−J11J22J34J43−J11J44J23J32+J11J23J34J42+J11J24J32J43−J11J33J24J42−J12J21J33J44+J12J21J34J43. | (3.8) |
Here,
J11=−r1X∗K,J22=−αr2X∗S∗cK,J33=−ξ−d2u,J44=−μv−λ1S∗,J12=−αX∗,J21=αr2S∗(1−S∗+I∗cK),J23=−αr2X∗S∗cK,J24=−λS∗J32=λV∗,J34=λS∗,J42=−λ1V∗,J43=κξ. | (3.9) |
According to the Routh-Hurwitz criterion, characteristic equation have roots with negative parts if
σ1>0,σ4>0,σ1σ2−σ3>0,σ1σ2σ3−σ23−σ21σ4>0. | (3.10) |
From Proposition 1 and from the above analysis, the following theorem is obtained.
Theorem 1. In the system (2.1),
(i) plant-pest free equilibrium E1 is always unstable,
(ii) pest-free equilibrium E2 is stable if αr2μvK<πvλ and unstable otherwise,
(iii) interior equilibrium E∗ exists if αr2μvK<πvλ, i.e., when E2 becomes unstable. E∗ is stable if the conditions in (3.10) are satisfied.
Now, we shall analyse the conditions for which E∗ enters into Hopf bifurcation as a model parameter varies over R. We consider the Hopf bifurcation as a function of the generic bifurcation parameter η∈R.
Let Ψ : (0,∞)→R be the following continuously differentiable function of η:
Ψ(η):=σ1(η)σ2(η)σ3(η)−σ23(η)−σ4(η)σ21(η) |
The assumptions for Hopf bifurcation to occur are the usual ones and these require that the spectrum σ(η)={ρ:D(ρ)=0} of the characteristic equation is such that
(A) There exists η∗∈(0,∞), at which a pair of complex eigenvalues ρ(η∗),ˉρ(η∗)∈σ(η) are such that
Reρ(η∗)=0, Imρ(η∗)=ω0>0, |
and the transversality condition
dReρ(η)dη|η∗≠0; |
(B) All other elements of σ(η) have negative real parts.
Thus we have the following theorem [28].
Theorem 2. The system (2.1) around the interior equilibrium E∗ enters into Hopf bifurcation at η=η∗∈(0,∞) if and only if
(i) Ψ(η∗)=0, and
(ii) σ31σ′2σ3(σ1−3σ3)>2(σ2σ21−2σ23)(σ′3σ21−σ′1σ23),
and all other eigenvalues have negative real parts, where ρ(η) is purely imaginary at η=η∗.
In the this section, we analyse the global stability of the coexisting equilibrium of system (2.1).
For the global stability, we choose the Lyapunov function as follows:
ψ(X,S,I,V)=12(c1X2+c2S2+c3I2+c4V2). |
Here, ci>0,i=1,2,3,4, are so chosen that ˙ψ is negative definite. Now, derivative of ψ along the solution of the equation ˙X(t)=JE∗X(t), where X(t)=(X(t),S(t),I(t),V(t))T, is as follows
˙ψ=c1X˙X+c2s˙S+c3I˙I+c4v˙V=c1(−r1X∗K)X2+c2(−αr2X∗S∗cK)S2−c3ξI2−c4(μv+λ1S∗)V2+(c2αr2S∗(1−S∗+I∗cK)−c1αX∗)XS+(c3λV∗−c2αr2X∗S∗cK)SI−(c2λS∗+c4λ1V∗)SV+(c4κξ+c3λS∗)IV |
Thus symmetric matrix corresponding to ˙ψ is given by
M=[mij]4×4=12[m11m1200m21m22(c3λV∗−c2αr2X∗S∗cK)−(c2λS∗+c4λ1V∗)0(c3λV∗−c2αr2X∗S∗cK)−2c3ξ(c4κξ+c3λS∗)0−(c2λS∗+c4λ1V∗)(c4κξ+c3λS∗)−2c4(μv+λ1S∗)]. | (4.1) |
Here,
m11=2c1(−r1X∗K),m12=m21=(c2αr2S∗(1−S∗+I∗cK)−c1αX∗,m22=2c2(−αr2X∗S∗cK). | (4.2) |
The positive equilibrium E∗ is locally asymptotically stable if ˙ψ is negative definite which implies the matrix M must be negative definite. But the matrix M will be negative definite if all of the principal minors of odd rank are negative and all of the principal minors of even rank are positive. This gives rise to the following four conditions:
(i)−2c1(r1X∗K)<0,(ii)4c1c2(−r1X∗K)(−αr2X∗S∗cK)−(c2αr2S∗(1−S∗+I∗cK)−c1αX∗)2>0,(iii)2c1(−r1X∗K)[4c3c2(αr2X∗S∗cK)ξ−(c3λV∗−c2αr2X∗S∗cK)2]+2c3ξ(c2αr2S∗(1−S∗+I∗cK)−c1αX∗)2<0,(iv)−2c4(μv+λ1S∗)×LHS of expression of inequality(iii)−(c4κξ+c3λS∗)×Minor corresponding toM43−(c2λS∗+c4λ1V∗)×Minor corresponding toM42>0. | (4.3) |
Clearly, first condition is automatically satisfied. Now, if we choose c2,c3 in such a way that
(c2αr2S∗(1−S∗+I∗cK)−c1αX∗)=0,and(c3λV∗−c2αr2S∗ks)=0, | (4.4) |
i.e.,
c2=c1αX∗αr2S∗(1−S∗+I∗cK)=c1αX∗2λS∗V∗,c3=c21λV∗αr2S∗ks, | (4.5) |
then the condition (ⅱ) and (ⅲ) are satisfied, and thus the condition (ⅳ) reduces to the following form:
16c1c2c3c4m11m22ξ(μv+λ1S∗)−m11m22(c4κξ+c3λS∗)2+2m11c3ξ(c2λS∗+c4λ1V∗)2>0. | (4.6) |
The above condition is satisfied for any suitable large value of c1 as first term in the above inequality is positive and other two terms are negative. The negative terms do not contain c1.
Thus, we have the following theorem for the global stability of E∗.
Theorem 3. The equilibrium point E∗(X∗,S∗,I∗,V∗) of the system (2.1) is locally asymptotically stable if the following condition holds:
16c1c2c3c4m11m22ξ(μv+λ1S∗)−m11m22(c4κξ+c3λS∗)2+2m11c3ξ(c2λS∗+c4λ1V∗)2>0, | (4.7) |
where m11 and m22 are given in (4.2) and c1,c2,c3, and c4 are given in (4.5).
To investigate the orbital stability of the Hopf-bifurcating periodic solution, ref30's condition has been followed [29]. According to ref30's sufficient condition, the supercritical and subcritical nature of the Hopf-bifurcating periodic solution is determined respectively by the positive and negative sign of real part of the number ϝ, where ϝ is defined by:
ϝ=−ul∂3fl∂xp∂xq∂xrvpvqˉvr+2ul∂2fl∂xp∂xqvp[(J−1E∗)qs]∂2fs∂xt∂xwvtˉvw+ul∂2fl∂xp∂xqˉvp[(JE∗−2iω0I)−1qs]∂2fs∂xt∂xwvtvw, | (5.1) |
where the repeated indices within each term imply a sum notation and all the derivatives of fl are evaluated at the equilibrium E∗. JE∗ is the variational matrix of the system (2.1) calculated at E∗. u=(u1,u2,u3,u4) and v=(v1,v2,v3,v4)T are the left and right eigenvectors respectively of E∗ with respect to eigenvalues iω0. So positivity of the real part of the above expression in parenthesis really indicates the orbital stability of the periodic solution arising out of Hopf bifurcation.
We rewrite our system (2.1) in the following form:
dxdt=f(x,t), | (5.2) |
where x=(X,S,I,V), f=(f1,f2,f3,f4)T, and fl, l=1,2,3,4 are right sides of system (2.1) i.e. f1=r1X(1−XK)−αXS etc. Now, all the second and third-order derivatives of fl(l=1,2,3,4) are as follows:
∂2f1∂X2=−2r1K,∂2f1∂j∂S=∂2f1∂S∂j=−α,∂2f2∂S2=−2αr2X∗cK,∂2f2∂j∂S=∂2f2∂S∂j=αr2−2αr2S∗cK∂2f2∂s∂i=∂2f2∂i∂s=−αr2X∗cK,∂3f2∂X∂S∂S=∂3f2∂s∂j∂s=∂3f2∂s∂S∂j=−2αr2cK,∂2f2∂X∂I=∂2f2∂I∂X=−αr2S∗cK,∂3f2∂j∂s∂i=∂3f2∂j∂i∂s=∂3f2∂s∂j∂i=∂3f2∂s∂i∂j=∂3f2∂i∂j∂s=∂3f2∂i∂s∂j=−αr2cK∂2f2∂s∂v=∂2f2∂v∂s=−λ∂2f3∂v∂s=∂2f3∂s∂v=λ,∂2f4∂s∂v=∂2f4∂v∂s−λ1. | (5.3) |
Now we calculate Mω0=(JE∗−2iω0I)−1
=1m[a11a12a11a11a21a22a23a24a31a32a33a34a41a42a43a43]. |
Here,
a11=−(αr2X∗S∗cK+2iω0){(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}+αr2X∗S∗cK{λV∗(μv+λ1S∗+2iω0)−λλ1S∗V∗}−λS∗{λκξV∗−λ1V∗(ξ+2iω0)}a12=αX∗{(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}a13=−αX∗{αr2X∗S∗cK(μv+λ1S∗+2iω0)+κξλS∗}a14=αX∗λS∗{αr2X∗S∗cK−(ξ+2iω0)}a21=−αr2S∗(1−S∗+I∗cK){(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}a22=−(r1X∗K+2iω0){(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}a23=(r1X∗K+2iω0){κξλS∗−αr2X∗S∗cK(μv+λ1S∗+2iω0)}a24=−λS∗(r1X∗K+2iω0){αr2X∗S∗cK−(ξ+2iω0)}a31=−αr2S∗(1−S∗+I∗cK){λV∗(μv+λ1S∗+2iω0)−λλ1S∗V∗}a32=−(r1X∗K+2iω0){λV∗(μv+λ1S∗+2iω0)−λλ1S∗V∗}a33=−(r1X∗K+2iω0){(αr2X∗S∗cK+2iω0)(μv+λ1S∗+2iω0)−λλ1S∗V∗}−αX∗αr2S∗(1−S∗+I∗cK)(μv+λ1S∗+2iω0)a34=(r1X∗K+2iω0){λ2S∗V∗−λS∗(αr2X∗S∗cK+2iω0)}+αX∗λS∗αr2S∗(1−S∗+I∗cK)a41=−αr2S∗(1−S∗+I∗cK){λV∗κξ−(ξ+2iω0)λ1V∗}a42=−(r1X∗K+2iω0){λV∗κξ−(ξ+2iω0)λ1V∗}a43=(r1X∗K+2iω0){αr2X∗S∗cKλ1V∗−κξ(αr2X∗S∗cK+2iω0)}+αX∗αr2S∗(1−S∗+I∗cK)κξa44=(r1X∗K+2iω0){(αr2X∗S∗cK+2iω0)(ξ+2iω0)−αr2X∗S∗cKλV∗}−αX∗αr2S∗(1−S∗+I∗cK)(ξ+2iω0) |
and
m=−(αr2X∗S∗cK+2iω0){(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}+αr2X∗S∗cK{λV∗(μv+λ1S∗+2iω0)−λλ1S∗V∗}−λS∗{λκξV∗−λ1V∗(ξ+2iω0)}+αX∗{(ξ+2iω0)(μv+λ1S∗+2iω0)−κξλS∗}. | (5.4) |
Now, if we put ω0=0 in the above expressions, we get the component of M=(JE∗)−1.
In the next section we find out the left eigenvector and right eigenvector of the variational matrix JE∗ with respect to the eigenvalue iω0, i.e., we calculate row vector u=(u1,u2,u3,u4) and column vector v=(v1,v2,v3,v4)T such that
uJE∗=iω0u,JE∗v=iω0v. | (5.5) |
Solving the first equation of (5.5), we find the left eigenvector u=(u1,u2,u3,u4) where
u1=m42m21S+iω0m242m21u2=m42{m11S−m42ω20}+iω0m42(m42m11+S)u3=m242m11m23−m42m43R+iω0m42{m42m23+m43Q}u4=m11m23m42m32−m42m33R+m42ω20Q+iω0m42{R+m33Q+m23m32}. |
Solving the second equations of (5.5), we find the right eigenvector v=η(v1,v2,v3,v4)T where
v1=m12m24T−iω0m12m224v2=m11m24T+m224ω20+iω0m24(T−m11m24)v3=m34m24R−m11m32m224−iω0m24(m34Q+m32m24)v4=m11m23m24m32+m24{m33R−ω20Q}+iω0m24(m23m32−m33Q−m24R). |
Now for uv=1 we get,
η=A−iω0BA2−ω20B2. | (5.6) |
Here,
A=m12m21m42m24UT+ω20m12m21m242m224+m42m24(m11U−m42ω20)(m11T+m24ω20)−ω20m42m24(m11m42+U)(T−m11m24)+m42m24(m11m23m42−m43R)(m34R−m11m32m24)+ω20m42m24(m23m42+m43Q)(m34Q++m32m24)m42m24{(m11m23m32)2−(m33R−ω20Q)2}−ω20m42m24{(m23m32)2−(R+m33Q)2},B=m12m21m242m24T−m12m21m42m224U+m42m24(T−m11m24)(m11U−m42ω20)+m42m24(m11m23m42−m43R)(m34Q−m32m24)+m42m24(m23m42+m43Q)(m34R−(m11m32m24)+m42m24(m11m32m23−m33R+ω20Q)(m32m23−m33Q+m24R)+m42m24(R+m33Q+m32m23)(m11m32m23+m33R−ω20Q), |
where,
Q=m11−m22,R=m12m21+m11m22+ω20,U=m32m43−m42m33,T=m24m33−m23m34. | (5.7) |
Writing the expression of (5.1) in detail, we have the following:
The first term:
−ul∂3fl∂xp∂xq∂xrvpvqˉvr=−u2∂3f2∂s∂j∂s(v22¯v1+2v1|v2|2)−u2∂3f2∂j∂s∂i(2v1v2ˉv3+2v1v3ˉv2+2v3v2ˉv1). | (5.8) |
The second term:
2ul∂2fl∂xp∂xqvp[(J−1E∗)qs]∂2fs∂xt∂xwvtˉvw | (5.9) |
=2[u1(∂2f1∂j2v1+∂2f1∂s∂jv2)+u2∂2f2∂s∂jv2][M11A1+M12A2+M13A3+M14A4]+2[u1∂2f1∂j∂sv1+u2(∂2f2∂j∂sv1+∂2f2∂i∂sv3+∂2f2∂v∂sv4)+u3∂2f3∂v∂sv4+u4∂2f4∂v∂sv4][M21A1+M22A2+M23A3+M24A4]+2u2(∂2f2∂j∂iv1+∂2f2∂s∂iv2)×[M31A1+M32A2+M33A3+M34A4]+2(u2∂2f2∂s∂vv2+u3∂2f3∂s∂vv2+u4∂2f4∂s∂vv2)×(M41A1+M42A2+M43A3+M44A4), | (5.10) |
The third term:
ul∂2fl∂xp∂xqˉvp[(JE∗−2iω0I)−1qs]∂2fs∂xt∂xwvtvw | (5.11) |
=[u1(∂2f1∂j2ˉv1+∂2f1∂s∂jˉv2)+u2∂2f2∂s∂jˉv2][Mω011B1+Mω012B2+Mω013B3+Mω014B4][u1∂2f1∂j∂sˉv1+u2(∂2f2∂j∂sˉv1+∂2f2∂i∂sˉv3+∂2f2∂v∂sˉv4)+u3∂2f3∂v∂sˉv4+u4∂2f4∂v∂sˉv4][Mω021B1+Mω022B2+Mω023B3+Mω024B4]+u2(∂2f2∂j∂iˉv1+∂2f2∂s∂iˉv2)[Mω031B1+Mω032B2+Mω033B3+Mω034B4]+(u2∂2f2∂s∂vˉv2+u3∂2f3∂s∂vˉv2+u4∂2f4∂s∂vˉv2)[Mω041B1+Mω042B2+Mω043B3+Mω044B4], | (5.12) |
where
A1=∂2f1∂j2|v1|2+∂2f1∂s∂j(v1ˉv2+v2ˉv1)A2=∂2f2∂s∂j(v1ˉv2+v2ˉv1)+∂2f2∂j∂i(v1ˉv3+v3ˉv1)+∂2f2∂s∂i(v3ˉv2+v2ˉv3)+∂2f2∂s∂j(v4ˉv2+v2ˉv4)A3=∂2f3∂s∂v(v4ˉv2+v2ˉv4)A4=∂2f4∂s∂v(v4ˉv2+v2ˉv4)B1=∂2f1∂j2(v1)2+2∂2f1∂s∂jv1v2B2=2∂2f2∂s∂jv1v2+2∂2f2∂j∂iv1v3+2∂2f2∂s∂iv2v3+2∂2f2∂s∂jv2v4B3=2∂2f3∂s∂vv2v4B4=2∂2f4∂s∂vv2v4 |
Putting the value of all second- and third-order derivatives of fl(l=1,2,3,4), u,v, and the components of matrix M and Mω0 in the first term, and placing the second term and third term in terms of the parameters of the model, we calculate the expression (5.1) and the sign of the real part of this expression. This in turn indicates the orbital stability of the limit cycle arising from Hopf bifurcation.
In this section, we have analyzed the dynamics of the model system using numerical simulations in MATLAB. The analytical results obtained in the previous sections are verified using numerical calculations. Results are plotted in the figures and discussed below.
In Figure 1, the numerical solution of the model is presented. As the value of the parameter is enhanced, oscillation in the solutions is increased. Solutions become periodic when the rate is higher (λ=0.00012).
Parameters | Short description | Value (unit) |
r1 | growth rate of crop biomass | 0.05 kg day−1 |
K | maximum crop biomass | 50 kg plant−1 |
α | crop consumption rate | 0.001 kg pest−1 day−1 |
λ | contact rate of pest with viruses | 0.00005 day−1 |
λ1 | reduction rate constant of virus | 0.00001 gm pest −1 day−1 |
when it attack pests | ||
ξ | mortality rate of infected pest | 0.1 plant−1 day−1 |
r2 | growth rate of susceptible pest | 8 day−1 |
μV | decay rate of virus | 0.1 gm day−1 |
c | a proportional constant | 6 |
πv | recruitment of biopesticides | 2.5 gm day−1 |
Figure 2 shows the global stability of interior equilibrium point E∗ for λ=0.00008. All phase portraits are converging to the same interior steady point (E∗(35.98,56.07,28.92,1031).
Figure 3 depicts the bifurcation diagram of the system. We have plotted the maximum and minimum values of the periodic solutions. When the bifurcation parameter λ crosses the critical value λ∗=0.0000965(approx.), the system bifurcates into periodic oscillations.
In Figure 4, we see the orbital stability of the bifurcating periodic solution for a fixed value of the parameter λ. We observe that when λ passes through the value 0.0000965 (approx.), the interior equilibrium E∗ bifurcates toward a periodic solution (see Figure 3). From this figure, we conclude that the bifurcating limit cycle is stable (supercritical).
Control of pest attacks is an important aspect in agriculture to obtain healthy crops as well as high yield. Mathematical modeling helps in identifying the parameters important for crop pest management. In this paper, we have considered a model for pest control using biological agents and observed the effects of biopesticide in controlling the pest. We have explored the global stability of the interior equilibrium point E∗. Applying Poore's criteria, we studied the stability of the limit cycle around the interior equilibrium point.
We have shown how the dynamics changes with the increase in the value of the parameter λ (the infection rate of the pest by the virus) of the system. The model reveals that infection can be sustained only above a threshold value of the consumption rate λ. On increasing the value of λ, the endemic equilibrium bifurcates toward a periodic solution (Theorem 3.2). Numerically, we have shown that system (2.1) is stable globally asymptotically when the consumption rate by the pest is below a threshold value and, after that value, the system is unstable for some higher value of this threshold value giving a stable periodic solution (Figure 2).
In conclusion, in this research, two important dynamical behaviors, namely the global stability of the endemic equilibrium and stability of Hopf bifurcating periodic solution, have been successfully analyzed analytically and numerically using a mathematical model for biological control of crop pests. This article established that the endemic equilibrium is globally stable when the consumption rate of the crop biomass by pests is lower. Also, a Hopf bifurcating periodic solution exits for a higher consumption rate and it is stable. The results will help us in proposing a proper control strategy for pest control in crop cultivation.
J.C., F.A.B.: Conceptualization; J.C., F.A.B.: Methodology; A.A.R., F.A.B.: Software; A.A.R., F.A.B.: Validation; A.A.R., J.C.: Formal analysis; A.A.R, F.A.B.: Investigation; A.A.R., J.C.: Writing-original draft preparation; J.C., F.A.B.: Writing-review and editing; F.A.B.: Supervision; A.A.R.: Project administration. All authors have read and agreed to the published version of the manuscript.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare no conflict of interest.
The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University, Abha, Kingdom of Saudi Arabia, for funding this work through Large Research Project under grant number RGP2/174/45.
[1] | S. Gupta, A. K. Dikshit, Biopesticides: An ecofriendly approach for pest control, J. Biopest, (2010), 186. |
[2] |
W. J. Lewis, J. C. Van Lenteren, S. C. Phatak, J. H. Tumlinson, A total system approach to sustainable pest management, Proc. Natl. Acad. Sci., 94, (1997), 12243–12248. https://doi.org/10.1073/pnas.94.23.1224 doi: 10.1073/pnas.94.23.1224
![]() |
[3] | M. L. Flint, R. Van den Bosch, Introduction to integrated pest management, Springer Science & Business Media, (2012). |
[4] | E. Beltrami, Mathematics for dynamic modeling, Academic press, (2014). |
[5] | R. M. May, Stability and complexity in model ecosystems, Princeton university press, (2019). |
[6] |
L. F. Cavalieri, H. Koçak, Chaos in biological control systems, J. Theoret. Biol., 169 (1994), 179–187. https://doi.org/10.1006/jtbi.1994.1139 doi: 10.1006/jtbi.1994.1139
![]() |
[7] |
W. L. Keith, R. H. Rand, Dynamics of a system exhibiting the global bifurcation of a limit cycle at infinity, Int. J. Non-Lin. Mech., 20 (1985), 325–338. https://doi.org/10.1016/0020-7462(85)90040-X doi: 10.1016/0020-7462(85)90040-X
![]() |
[8] | S. Sastry, Nonlinear systems: Analysis, stability, and control, Springer Science, Business Media, 10 (2013). |
[9] | R. Seydel, Practical bifurcation and stability analysis, Springer Science & Business Media, (2009). |
[10] |
Z. He, X. Lai, Bifurcation and chaotic behavior of a discrete-time predator–prey system, Nonlinear Anal. Real. World Appl., 12 (2019), 403–417. https://doi.org/10.1016/j.nonrwa.2010.06.026 doi: 10.1016/j.nonrwa.2010.06.026
![]() |
[11] | S. H. Strogatz, Nonlinear dynamics and chaos with student solutions manual: With applications to physics, biology, chemistry, and engineering, CRC press, (2018). |
[12] |
V. Kumar, J. Dhar, H. S. Bhatti, Stability and Hopf bifurcation dynamics of a food chain system: plant–pest–natural enemy with dual gestation delay as a biological control strategy, Model. Earth Syst. Environ., 4 (2018), 881–889. https://doi.org/10.1007/s40808-018-0417-1 doi: 10.1007/s40808-018-0417-1
![]() |
[13] |
F. A. Basir, A multi-delay model for pest control with awareness induced interventions—Hopf bifurcation and optimal control analysis, Int. J. Biomath., 13 (2020), 2050047. https://doi.org/10.1142/S1793524520500473 doi: 10.1142/S1793524520500473
![]() |
[14] |
T. Abraha, F. Al Basir, L. L. Obsu, D. F. M. Torres, Farming awareness based optimum interventions for crop pest control, Math. Biosci. Eng., 18 (2021), 5364–5391. https://doi.org/10.3934/mbe.2021272 doi: 10.3934/mbe.2021272
![]() |
[15] | W. Costello, H. Taylor, Mathematical models of the sterile male technique of insect control, in: Mathematical Analysis of Decision Problems in Ecology, Springer, Berlin, Heidelberg, (1975), 318–359. https://doi.org/10.1007/978-3-642-80924-8_12 |
[16] |
T. L. Vincent, Pest management programs via optimal control theory, Biometrics, 31 (1975), 1–10. https://doi.org/10.2307/2529704 doi: 10.2307/2529704
![]() |
[17] |
Y. Liu, Y. Yang, B. Wang, Entomopathogenic fungi Beauveria bassiana and Metarhizium anisopliae play roles of maize (Zea mays) growth promoter, Sci. Rep., 12 (2022), 15706. https://doi.org/10.1038/s41598-022-19899-7 doi: 10.1038/s41598-022-19899-7
![]() |
[18] |
F. A. Basir, S. Samanta, P. K. Tiwari, Bistability, generalized and zero-hopf bifurcations in a pest control model with farming awareness, J. Biol. Syst., 31 (2023), 115–140. https://doi.org/10.1142/S0218339023500079 doi: 10.1142/S0218339023500079
![]() |
[19] |
G. Seo, G. S. Wolkowicz, Pest control by generalist parasitoids: A bifurcation theory approach. Discrete Cont. Dyn. S., 31 (2020), 3157–3187. https://doi.org/10.3934/dcdss.2020163 doi: 10.3934/dcdss.2020163
![]() |
[20] | D. K. Bhattacharya, S. Karan, On bionomic model of integrated pest management of a single pest population, J. Differ. Equat. Dyn. Syst., 12 (2004), 301–330. |
[21] |
S. Ghosh, D. K. Bhattacharyya, Optimization in microbial pest control: An integrated approach, Appl. Math. Model., 34 (2010), 1382–1395. https://doi.org/10.1016/j.apm.2009.08.026 doi: 10.1016/j.apm.2009.08.026
![]() |
[22] | F. A. Basir, A. Banerjee, S. Ray, Role of farming awareness in crop pest management—a mathematical model, J. Theoret. Biol., 461 (2019), 59–67. |
[23] | E. Kurstak, Microbial and Viral Pesticide, Marcel and Dekker, Inc., New York, Bessel, (1982). |
[24] |
S. Bhattacharyya, D. K. Bhattacharyya, An improved integrated pest management model under 2-control parameters (sterile male and pesticide), Math. Biosci., 209, (2007), 256–281. https://doi.org/10.1016/j.mbs.2006.08.003 doi: 10.1016/j.mbs.2006.08.003
![]() |
[25] | J. Chowdhury, F. Al Basir, J. Pal, P. K. Roy, Pest control for Jatropha curcas plant through viral disease: a mathematical approach, Nonlinear Stud., 23 (2016), 517–532. |
[26] |
T. Abraha, F. A. Basir, L. L. Obsu, D. F. M. Torres, Pest control using farming awareness: Impact of time delays and optimal use of biopesticides, Chaos Soliton. Fract., 146 (2021), 110869. https://doi.org/10.1016/j.chaos.2021.110869 doi: 10.1016/j.chaos.2021.110869
![]() |
[27] |
S. Ghosh, S. Bhattacharyya, D.K. Bhattacharyya, The Role of Viral infection in Pest Control: A Mathematical Study, Bull. Math. Biol., 69 (2007), 2649–2691. https://doi.org/10.1007/s11538-007-9235-8 doi: 10.1007/s11538-007-9235-8
![]() |
[28] |
J. Chowdhury, F. A. Basir, Y. Takeuchi, M. Ghosh, P. K. Roy, A mathematical model for pest management in Jatropha curcas with integrated pesticides–an optimal control approach, Ecol. Complex., 37 (2019), 24–31. https://doi.org/10.1016/j.ecocom.2018.12.004 doi: 10.1016/j.ecocom.2018.12.004
![]() |
[29] |
A. B. Poore, On the theory and application of the Hopf-Friedrichs bifurcation theory, Arch. Rat. Mech. Anal., 60 (1976), 371–393. https://doi.org/10.1007/BF00248886 doi: 10.1007/BF00248886
![]() |
Parameters | Short description | Value (unit) |
r1 | growth rate of crop biomass | 0.05 kg day−1 |
K | maximum crop biomass | 50 kg plant−1 |
α | crop consumption rate | 0.001 kg pest−1 day−1 |
λ | contact rate of pest with viruses | 0.00005 day−1 |
λ1 | reduction rate constant of virus | 0.00001 gm pest −1 day−1 |
when it attack pests | ||
ξ | mortality rate of infected pest | 0.1 plant−1 day−1 |
r2 | growth rate of susceptible pest | 8 day−1 |
μV | decay rate of virus | 0.1 gm day−1 |
c | a proportional constant | 6 |
πv | recruitment of biopesticides | 2.5 gm day−1 |
Parameters | Short description | Value (unit) |
r1 | growth rate of crop biomass | 0.05 kg day−1 |
K | maximum crop biomass | 50 kg plant−1 |
α | crop consumption rate | 0.001 kg pest−1 day−1 |
λ | contact rate of pest with viruses | 0.00005 day−1 |
λ1 | reduction rate constant of virus | 0.00001 gm pest −1 day−1 |
when it attack pests | ||
ξ | mortality rate of infected pest | 0.1 plant−1 day−1 |
r2 | growth rate of susceptible pest | 8 day−1 |
μV | decay rate of virus | 0.1 gm day−1 |
c | a proportional constant | 6 |
πv | recruitment of biopesticides | 2.5 gm day−1 |