
The current research presents a predator-prey model that incorporates both a Gilpin-Ayala growth function and a Holling type Ⅲ functional response. Two Lyapunov functions are established to confirm the global asymptotic stability of the positive equilibrium P∗ and the predator extinction equilibrium Pk. Considering ecological protection and commercial incentives, we also incorporated a weighted harvesting strategy and pulse control into the model. We investigated intricate dynamical problems instigated by the weighting harvesting and pulse effects, and affirmed the existence and local asymptotic stability of both predator-extinction periodic solution and positive order-1 periodic solution. In the end, a suite of numerical simulations were carried out using MATLAB, aiming to corroborate the theoretical findings and deliver conclusions rooted in a biological context.
Citation: Xiaohuan Yu, Mingzhan Huang. Dynamics of a Gilpin-Ayala predator-prey system with state feedback weighted harvest strategy[J]. AIMS Mathematics, 2023, 8(11): 26968-26990. doi: 10.3934/math.20231380
[1] | Ilyasse Lamrani, Imad El Harraki, M. A. Aziz-Alaoui, Fatima-Zahrae El Alaoui . Feedback stabilization for prey predator general model with diffusion via multiplicative controls. AIMS Mathematics, 2023, 8(1): 2360-2385. doi: 10.3934/math.2023122 |
[2] | Shanshan Yu, Jiang Liu, Xiaojie Lin . Multiple positive periodic solutions of a Gause-type predator-prey model with Allee effect and functional responses. AIMS Mathematics, 2020, 5(6): 6135-6148. doi: 10.3934/math.2020394 |
[3] | Heping Jiang . Complex dynamics induced by harvesting rate and delay in a diffusive Leslie-Gower predator-prey model. AIMS Mathematics, 2023, 8(9): 20718-20730. doi: 10.3934/math.20231056 |
[4] | 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 |
[5] | Jagdev Singh, Behzad Ghanbari, Ved Prakash Dubey, Devendra Kumar, Kottakkaran Sooppy Nisar . Fractional dynamics and computational analysis of food chain model with disease in intermediate predator. AIMS Mathematics, 2024, 9(7): 17089-17121. doi: 10.3934/math.2024830 |
[6] | Chuanfu Chai, Yuanfu Shao, Yaping Wang . Analysis of a Holling-type IV stochastic prey-predator system with anti-predatory behavior and Lévy noise. AIMS Mathematics, 2023, 8(9): 21033-21054. doi: 10.3934/math.20231071 |
[7] | Anel Esquivel-Navarrete, Jorge Sanchez-Ortiz, Gabriel Catalan-Angeles, Martin P. Arciga-Alejandre . Tritrophic fractional model with Holling III functional response. AIMS Mathematics, 2024, 9(6): 15937-15948. doi: 10.3934/math.2024771 |
[8] | Mianjian Ruan, Chang Li, Xianyi Li . Codimension two 1:1 strong resonance bifurcation in a discrete predator-prey model with Holling Ⅳ functional response. AIMS Mathematics, 2022, 7(2): 3150-3168. doi: 10.3934/math.2022174 |
[9] | Chengchong Lu, Xinxin Liu, Zhicheng Li . The dynamics and harvesting strategies of a predator-prey system with Allee effect on prey. AIMS Mathematics, 2023, 8(12): 28897-28925. doi: 10.3934/math.20231481 |
[10] | Chuangliang Qin, Jinji Du, Yuanxian Hui . Dynamical behavior of a stochastic predator-prey model with Holling-type III functional response and infectious predator. AIMS Mathematics, 2022, 7(5): 7403-7418. doi: 10.3934/math.2022413 |
The current research presents a predator-prey model that incorporates both a Gilpin-Ayala growth function and a Holling type Ⅲ functional response. Two Lyapunov functions are established to confirm the global asymptotic stability of the positive equilibrium P∗ and the predator extinction equilibrium Pk. Considering ecological protection and commercial incentives, we also incorporated a weighted harvesting strategy and pulse control into the model. We investigated intricate dynamical problems instigated by the weighting harvesting and pulse effects, and affirmed the existence and local asymptotic stability of both predator-extinction periodic solution and positive order-1 periodic solution. In the end, a suite of numerical simulations were carried out using MATLAB, aiming to corroborate the theoretical findings and deliver conclusions rooted in a biological context.
Biological populations in nature exhibit intricate interactions, including competition, cooperation and predation. The predator-prey relationship is one of the most prevalent ecological relationships in the natural world. Both predator and prey play pivotal roles as vital biological resources, impacting human life and economic growth. To preserve the renewal potential of these biological resources, rational development and scientific management are imperative[12,16,24]. Employing limited renewable resources for sustainable development and utilization has emerged as a shared interest among economic management scientists, mathematicians and ecologists [4,5].
Italian mathematician V. Volterra introduced the Lotka-Volterra model in 1926 to depict variations in fish population. This model still serves as a foundational reference for predator-prey interaction models today [18,19,28,34]. However, the classic Lotka-Volterra model overlooks the limited capacity for predation in its portrayal of predator-prey relationships, thereby oversimplifying the functional response of the predator to the prey. Holling addressed this oversight in 1965 by introducing three distinct functional response functions [11]. Subsequent studies, such as that by M. Zhang and L. Chen [32], delved into a state feedback impulsive controlled predator-prey system with the Holling-Ⅱ functional response. They examined its equilibrium and presented the state feedback model, which assumes the harvest of predators and supplementation of preys as impulsive disturbances. This model aims to find the dynamical balance between prey and predator populations. Further research by M. Huang et al. [13,14,15] investigated a class of time dependent impulsive switching systems and discussed a series of population control strategies. Nevertheless, state feedback impulsive control proves to be more appropriate for studying predator-prey systems [29,33]. In contrast to time dependent harvesting strategies, state feedback harvesting strategies consider the current status of species[25,30], and reduce potential adverse effects on species' sustainability. The weighted sum of harvesting methods, which take into account the states of both predator and prey populations [7,24], promote the sustained development of natural populations.
Biomathematical researchers have previously used logistic growth models to characterize predator-prey relationships. However, such models tend to depict population growth in a linear manner, which is not always accurate. To rectify this shortcoming, Gilpin-Ayala introduced the exponential parameter α to the classical logistic population growth term. In 1973, Ayala and Gilpin improved the logistic growth model, proposing the following model [1]:
dAdt=A(r−KAα). |
By introducing a parameter α∈(0,1), they transformed the model into a nonlinear system, thereby increasing its analytical complexity [22]. Nevertheless, this exponential parameter α overcomes the limitations of the classical logistic population growth theory, specifically the assumption that each addition to the population reduces the population growth rate by a constant. This refined model offers a more nuanced and accurate representation of various population dynamics. Building on this idea, Jiang et al. [17,26,27] further developed this concept by allowing the Gilpin-Ayala parameter to vary with state under Markov switching and pulse interference.
Gonzˇslez-Olivares and Rojas-Palma [9] furthered this area of study by proposing a model incorporating the Holling Ⅲ functional response and the Allee effect:
{dAdt=r(1−AK)(A−m)A−τ1A2BA2+β,dBdt=(τ1A2A2+β−ω)B. |
Further research on this model was presented in [30] where the model construction incorporated nonlinear impulse disturbances. Both the predation mortality and predator release rates depend on their densities, adding realism to the model.
The Gilpin-Ayala growth model provides a more precise representation of population density changes. By introducing the exponential parameter, it overcomes the inherent limitations of the classic logistic population growth theory. This theory traditionally posits that each addition to the population uniformly reduces the growth rate. In reality, the Gilpin-Ayala model offers a more insightful description of the dynamic behaviors of population organisms and stands as one of the most critical numerical models in biological research [2]. Moreover, compared to the initial two Holling functional response functions [32], the Holling Ⅲ type [8,21] more authentically mirrors consumption rate variations at different resource densities. Based on the above literature analysis and research from [20], this paper formulates a predator-prey model that incorporates the Gilpin-Ayala growth, the Holling Ⅲ type functional response and state feedback impulsive control.
This paper is organized as follows: In Section 2 we develop a Gilpin-Ayala predator-prey system with a discontinuous weighted harvest strategy and provide the lemmas and definitions to be used in subsequent sections. Section 3 presents proof of positivity and boundedness for the uncontrolled system, as well as the global stability of equilibrium points. We also discuss the existence and stability for the predator extinction periodic solution and positive order-1 periodic solution for the controlled system. In Section 4 we provide some numerical simulations to validate these theoretical results, and, finally, a brief conclusion is given in Section 5.
In this paper, we consider a sufficiently large prey population and thus neglect the Allee effect, such as in [3,9,20]. Consequently, we introduce a predator-prey model featuring both Gilpin-Ayala growth and Holling Ⅲ functional response as follows:
{dAdt=rA[1−(AK)α]−τ1A2BA2+β≐Af1(A,B),dBdt=(τ2A2A2+β−ω)B≐Bf2(A,B), | (2.1) |
where A(t) and B(t) denote the densities of prey and predator at time t, respectively. The variable r stands for the intrinsic growth rate of the prey population, while K represents the maximum environmental carrying capacity of the prey population in the absence of predators and harvesting. The symbol α is a positive parameter to modify the classical logistic model and represents a nonlinear measure of interspecific interference. The term τ1A2A2+β denotes the functional response of the predator, commonly referred to as the Holling type Ⅲ response function. The variables τ1, τ2, β and ω correspond to the predator capture rate, the conversion coefficient specifying the ratio of newborn predators to captured prey, the half-saturation constant, and the death rate of the predator, respectively.
Considering commercial interests and the sustainability of biological resources, reasonable harvesting is permitted. The authors in [30] focused only on monitoring the predator population state to trigger pulse harvesting, examining a nonlinear ecological control model with a complex discrete map. Similar harvesting mechanisms have been very common in previous studies [1,17,22,26,27,30]. In [24], a different state dependent harvesting method was proposed that permitted the harvesting only when the combined densities of both prey and predator, weighted by specific factors, reached a predefined threshold. This approach not only ensures the sustainable development of the two biological resources but also safeguards commercial interests. Drawing on their approach, we also consider a weighted state pulse harvest for the above system (2.1). Denote the predefined harvesting threshold for the weighted sum of the densities of prey and predator as I. The intensity of this harvesting is measured by h and the prey and predator populations are assigned weights of μ and 1−μ, respectively. The model driven by this weighted harvesting strategy can be expressed as follows:
{dAdt=rA[1−(AK)α]−τ1A2BA2+βdBdt=(τ2A2A2+β−ω)B} μA+(1−μ)B<I,A(t+)=A(t)−c1hA(t)B(t+)=B(t)−c2hB(t)+Λ} μA+(1−μ)B=I, | (2.2) |
where c1 and c2 represent capture rates, while Λ describes the constant number of predator released at time t.
The practice of harvesting two types of populations simultaneously and releasing a certain number of predators in model (2.2) is also in line with many practical application scenarios. For example, in fishing operations, when the combined count of phytoplankton or shrimp (representing the prey) and fish (acting as the predator) hits a certain limit, there's an initiative to harvest the fish. However, to optimize salvage costs and uphold ecological equilibrium, phytoplankton is simultaneously harvested. This not only boosts the economic returns, but also enhances fish yield per unit area. Subsequently, to guarantee the yield and maintain a balanced ecosystem, a specific quantity of fish fry is introduced into the environment. This approach mirrors the prevalent fishery practice known as "rotational catch and stocking."
The foundational concepts and theories related to state feedback pulse dynamic systems are important for our subsequent discussion. Descriptions of such dynamic systems, along with definitions of pulse set, phase set, successor function, order k-periodic solution and its stability, are commonly found in various literature. Given this prevalence, we will exclude certain rudimentary explanations here (interested readers can refer to [6]). Instead, we'll focus on detailing the definition for a distinct type of successor function and the criteria to determine the order-k periodic solution.
Definition 1. (Successor function[25,31]) Assume that the pulse set and the phase set of a state feedback pulse dynamic system are two lines denoted by ˉM and ˉN, respectively (see Figure 1). Let O0 be the intersection point of ˉN and the A-axis. For any point V1∈ˉN, we denote the trajectory that originates from V1 by Γ1(V1). If the trajectory Γ1(V1) intersects with the phase set ˉN and pulse set ˉM at points V2 and V3, and point V3 is mapped to a point V4∈ˉN through pulse effect, then S1(V2)=d(V4,O0)−d(V2,O0) is called the type-1 successor function of point V1, while S2(V1)=d(V4,O0)−d(V1,O0) is called the type-2 successor function of point V1, where the function d(⋅,⋅) represents the directed distance between two points.
Lemma 1. (Lasalle invariance principle) Consider the following system of differential equations
{dAdt=P(A,B),dBdt=Q(A,B), |
where (A,B)∈G⊂R2 with G is a bounded set. P(A,B) and Q(A,B) are continuous and satisfy the Lipschitz condition. Given a scalar function V(A,B) that is continuous over ˉG and possesses a continuous first-order partial derivative, V(A,B) is referred to as a Lyapunov function on G if
dV(A,B)dt=∂V∂AdAdt+∂V∂BdBdt≤0. |
Define
S={(A0,B0)∈ˉG∣dV(A0,B0)dt=0}. |
Assuming M⊂S is the largest invariant subset within S, and given a solution x(t,t0,(A0,B0))⊂G of the system, it follows that x(t,,t0,(A0,B0))→M as t→∞.
Lemma 2. (Analogue of the Poincare criterion[10,23,25]) The T-periodic solution ˜ρL(t)=(˜A(t),˜B(t)) of the system
{dAdt=P(A,B),dBdt=Q(A,B),ifϕ(A,B)≠0,ΔA=a(A,B),ΔB=b(A,B),ifϕ(A,B)=0, |
is orbitally asymptotically stable and exhibits asymptotic phase behavior if the Floquet multiplier μ2 satisfies |μ2|<1. Here,
μ2=n∏k=1Δkexp[∫T0(∂P∂A+∂Q∂B)(˜A(t),˜B(t))dt] |
and
Δk=P+(∂b∂B∂ϕ∂A−∂b∂A∂ϕ∂B+∂ϕ∂A)+Q+(∂a∂A∂ϕ∂B−∂a∂B∂ϕ∂A+∂ϕ∂B)P∂ϕ∂A+Q∂ϕ∂B, |
where P+=P(˜A(τ+k),˜N(τ+k)) and Q+=Q(˜A(τ+k),˜N(τ+k)). The values of P,Q,∂a∂A,∂a∂B,∂b∂A,∂b∂B,∂ϕ∂A,∂ϕ∂B are computed at the point (˜A(τk),˜B(τk)).
Let's suppose that A(0)=A0>0 and B(0)=B0>0 are positive initial conditions. Initially, we assess the system (2.1) with these conditions for the positivity and boundedness of the solutions. Considering that the right-hand sides of (2.1) are smooth functions with respect to both A and B and dAdt|A=0=dBdt|B=0=0, the state space for the system (2.1) is the positive quadrant {(A,B):A>0,B>0}, which is an invariant set. Consequently, the solutions of the system (2.1) are positive. Next, we verify the boundedness of the solutions for the system (2.1).
Undoubtedly,
dAdt=rA[1−(AK)α]−τ1A2BA2+β≤rA[1−(AK)α]. |
Let
rA[1−(AK)α]≐ˉf(A), |
where ˉf(A)≥0 for A∈(0,K], and ˉf(A)<0 for A>K. Thus, we obtain A(t)≤max{A(0),K}.
Subsequently, we denote N:=τ2τ1A+B, then we have
dNdt=τ2τ1dAdt+dBdt=−ωN+τ2τ1[ω+r−r(AK)α]A:=−ωN+˜f(A). |
Through direct calculation, we can deduce that the maximum point of ˜f(A) is Am=Kα√ω+rr(α+1). Denote the maximum value by Mm=˜f(Am), we have
dNdt≤−ωN+Mm |
and
N(t)≤max{τ2τ1A(0)+B(0),Mmω}. |
Define
Ω0:={(A,B)|0≤A≤K,0≤τ2τ1A+B≤Mmω}, |
and we can easily know that Ω0 is a positive invariant set for the system (2.1).
Model (2.1) always has two equilibria: a saddle point Po(0,0) and a boundary equilibrium Pk(K,0). Moreover, a positive equilibrium P∗(A∗,B∗) exists when τ2K2K2+β>ω, where
A∗=√ωβτ2−ω,B∗=r(1−(A∗K)α)τ2β(τ2−ω)τ1A∗. | (3.1) |
To examine the local stability of the equilibrium points Pk(K,0) and P∗(A∗,B∗), we first determine the Jacobian matrix for the system (2.1) at any point (A,B). It is given by
J=(r−r(α+1)(AK)α−2τ1βAB(A2+β)2−τ1A2A2+β2τ2βAB(A2+β)2τ2A2A2+β−ω). | (3.2) |
The stability of the equilibrium points will be determined by the eigenvalues of the Jacobian matrix calculated at the respective equilibrium points.
By a simple calculation, the trivial equilibrium point Po(0,0) is unstable as long as it exists. Substituting (A,B)=(K,0) into (3.2), we get the characteristic values λk1=−rα, λk2=τ2K2K2+β−ω. Clearly, if τ2K2K2+β<ω, then P∗ does not exist and Pk(K,0) is locally stable; if τ2K2K2+β>ω, i.e., P∗ exists, then Pk is unstable and both Pk and Po are saddle points.
From the above discussion, both Po(0,0) and Pk(K,0) are unstable when P∗ exists. To investigate the local asymptotic stability of the positive equilibrium point P∗(A∗,B∗), we calculate
JP∗=(r−r(α+1)(AK)α−2τ1βAB(A2+β)2−τ1A2A2+β2τ2βAB(A2+β)2τ2A2A2+β−ω), | (3.3) |
from which we obtain
det(P∗)=2ωr(1−ωτ2)(1−(A∗K)α)>0. |
The trace of P∗ is given by
tr(P∗)=r[1−(α+1)(A∗K)α−2τ2(τ2−ω)(1−(A∗K)α)]=rτ2[2ω−τ2+(A∗K)α((1−α)τ2−2ω)]=rτ2[2ω−τ2+K−α(ωβτ2−ω)α2((1−α)τ2−2ω)]. |
This implies that P∗ is locally asymptotically stable if the following condition holds
2ω−τ2+K−α(ωβτ2−ω)α2((1−α)τ2−2ω)<0,whereω≪τ2. |
In order to illustrate the global dynamic behavior of Pk(K,0) and P∗(A∗,B∗), we construct two Lyapunov functions in the following two theorems.
Theorem 1. The boundary equilibrium point Pk(K,0) of system (2.1) is globally asymptotically stable when τ2K2K2+β<ω.
Proof. Consider a function V(A,B) defined by
V(A,B)=1α(Aα−Kα)−KαlnAK+B. |
By taking the derivative of system (2.1), it can be obtained that
dVdt=(Aα−1−KαA)dAdt+dBdt=−rKα(Aα−Kα)2+τ1AB(Aα−Kα)+B[(τ2−ω)A2−βω]β+A2. |
Since τ2K2K2+β<ω and A≤K, we have (τ2−ω)K2−βω<0 and (τ2−ω)A2−βω<0. This leads to the conclusion that dVdt≤0. Moreover, dVdt=0 if, and only if, (A,B)=(K,0). By the Lasalle invariance principle (Lemma 1), the boundary equilibrium point Pk(K,0) of system (2.1) is globally asymptotically stable. The proof is completed.
The preceding discussion establishes that a positive equilibrium P∗ exists if τ2K2K2+β>ω. To assess the global stability of P∗, we let
F(A):=rτ1(1−(AK)α)(A2+β)A, |
ξ1(A):=A2−β+(AK)α[β(1−α)−A2(1+α)]. |
Theorem 2. The positive equilibrium point P∗(A∗,B∗) of system (2.1) is globally asymptotically stable when τ2K2K2+β>ω, if one of the following conditions holds:
(Ⅰ) ξ1(A)=0 has no or only one positive root.
(Ⅱ) ξ1(A)=0 has two positive roots A1 and A2 (where A1<A2). If A∗∈(0,A′1), then F(A)>B∗ for all A∈(0,A∗) and F(A)<B∗ for all A∈(A∗,K) holds; if A∗∈(A′2,K), then F(A)>B∗ for all A∈(0,A∗) and F(A)<B∗ for all A∈(A∗,K) holds, where A′1,A′2 are two points that must exist in this case such that F(A′1)=F(A2) and F(A′2)=F(A1) with the ordering 0<A′1<A1<A2<A′2<K.
Proof. Let us define a function
V(u,v)=∫AA∗(τ2−ω)u2−βωτ2u2du+τ1τ2∫BB∗v−B∗vdv, |
then we have
dVdt=(τ2−ω)A2−βωτ2A2dAdt+τ1τ2B−B∗BdBdt=(τ2−ω)A2−βωτ2A2{rA[1−(AK)α]−τ1A2BA2+β}+τ1τ2B−B∗B[(τ2A2A2+β−ω)B]=τ1τ2(τ2−ω)A2−βωβ+A2[rτ1(1−(AK)α)β+A2A−B∗]=τ1τ2(τ2A2β+A2−ω)(F(A)−B∗). |
Clearly, dVdt=0 if, and only if, (A,B)=(A∗,B∗).
Next, we consider the function
F(A):=rτ1(1−(AK)α)(A2+β)A, |
and compute its derivative with respect to A, we have
F′(A)=rτ1A2{A2−β+(AK)α[β(1−α)−A2(1+α)]}. |
Let
ξ1(A):=A2−β+(AK)α[β(1−α)−A2(1+α)], |
then it follows, by differentiate with respect to A, that
ξ′1(A)=2A+β(1−α)αAα−1kα−3α(1+α)Aα+1Kα. |
In the special case of α=1, we obtain the maximum point ξ1max(A)=K227−β of ξ1(A)=0. For the roots of ξ1(A)=0 we note that: (i) If K227<β, then there is no root. (ii) If K227=β, then there is a unique root A=Ac satisfying √β<Ac<K. (iii) If K227>β, then there are two roots A1,A2 satisfying √β<A1<A2<K.
For the case α∈(0,1), when β(1−α)−A2(1+α)=0, i.e., A=√β(1−α)1+α, we have ξ1(A)=A2−β<0. For A≠√β(1−α)1+α and ξ1(A)=0, we have
(AK)α=A2−βA2(1+α)−β(1−α):=ξ2(A). |
When 0<α<1, we observe that ξ2(0)=11−α>1,limA→{√1−α1+αβ}−ξ2(A)=+∞, limA→{√1−α1+αβ}+ξ2(A)=−∞,0<ξ2(√β)=0,ξ2(K)<1 and y1(A)=ξ2(A) is strictly monotonically increasing on intervals [0,√1−α1+αβ) and (√1−α1+αβ,K]. While y2(A)=(AK)α is strictly monotonically increasing on interval [0,K] with y2(0)=0 and y2(K)=1. The graphs of the functions y2(A)=ξ2(A) and y1(A)=(AK)α either have no intersection points, or they intersect at a single point Ac such that √β<Ac<K, or they intersect at two points, A1 and A2, where √β<A1<A2<K.
Using an analogous analysis for α>1, the distribution of the roots for ξ1(A)=0 exhibits a similar behavior. Therefore, for any α>0, there are three cases for the root of the function F′(A)=0. Let's summarize these aspects in two cases:
(I) For the cases (i) and (ii), we obtain that F(A)−B∗>0 when A<A∗ and F(A)−B∗<0 when A>A∗. This implies that dVdt≤0 for all (A,B)∈Ω0, and dVdt=0 if, and only if, (A,B)=(A∗,B∗) (see Figure 2(a) and (b)).
(II) For the case (iii), we identify two positive points, A1 and A2, such that F′(Ai)=0 for i=1,2. Additionally, two distinct points, A′1 and A′2, must exist satisfying F(A′1)=F(A2) and F(A′2)=F(A1), with the ordering 0<A′1<A1<A2<A′2<K (see Figure 2(c)). Based on this, we deduce the following:
(a) F(A)>B∗ for A∈(0,A∗) and F(A)<B∗ for A∈(A∗,K) when A∗∈(0,A′1);
(b) F(A)>B∗ for A∈(0,A∗) and F(A)<B∗ for A∈(A∗,K) when A∗∈(A′2,K).
Thus, we can also claim that dVdt≤0 for all (A,B)∈Ω0, and dVdt=0 if, and only if, (A,B)=(A∗,B∗) According to Lasalle invariance principle (Lemma 1), we affirm that the positive equilibrium point P∗(A∗,B∗) of system (2.1) is globally asymptotically stable if one of the conditions listed in Theorem 2 holds. The proof is completed.
To discuss the dynamic behavior of system (2.2), we first define the impulse set and phase set as follows:
ˉM={(A,B)|0≤A≤K,μA+(1−μ)B=I}, | (3.4) |
ˉN={(A,B)|μA1−c1h+(1−μ)B1−c2h=I+(1−μ)Λ1−c2h}. | (3.5) |
A trajectory of system (2.2) that is tangent to the phase set at (A,B)∈ˉN must satisfy the following conditions:
{B=[I+(1−μ)Λ1−c2h−μA1−c1h](1−c2h1−μ),dBdA=(τ2A2A2+β−ω)BrA[1−(AK)α]−τ1A2BA2+β=−μ(1−c2h)(1−c1h)(1−μ). | (3.6) |
Next, we will concentrate on the dynamic behavior of both predator-extinction periodic solution and positive periodic solution of the system (2.2).
For Λ=0, we obtain that B≡0 if B(0)=0. At this point system (2.2) can be transformed into the following form:
{dAdt=rA[1−(AK)α]B=0} μA<I,ΔA(t)=−c1hA(t)ΔB(t)=0} μA=I. | (3.7) |
Assuming that ˉN intersects the A-axis at the point L(A0,0), we denote A=˜A(t) as a solution of the equation A′(t)=rA(1−(A/K)α) with the initial condition A(0)=A0. We denote a solution of the system (3.7) as ˜ρL(t)=(˜A(t,A0),0), where
˜A(t,A0)=e−rt(A−α0−e−αrtKα)−1α. | (3.8) |
Let ˜T=∫I/μA0Kα/rA(Kα−Aα)dA=˜T(A0), which satisfies ˜A(˜T)=I/μ and ˜A(˜T+)=A0. This illustrates that A(t)=˜A(t,A0) is a periodic solution of (3.7) with period ˜T.
Theorem Under the conditions Λ=0 and I≤μK, if c2>˜c2, then the system (2.2) admits an orbitally asymptotically stable predator-extinction periodic solution ˜ρL(t)=(˜A(t,A0),0). Here, ˜c2 is defined as
˜c2:=1h(1−(Kαμα−Iα)exp(ln(1−c1h)−∫˜T0˜Af1A(˜A,0)+f2(˜A)dt)(1−c1h)(Kαμα−(1−c1h)αIα)). |
Proof. Assume I>μK, we notice that the trajectory of system (3.7) will not reach A=I/μ for any initial value (A(t0),0). Therefore, there is no periodic solution in the system (3.7) and the predator-extinction equilibrium (K,0) is asymptotically stable. Conversely, if I/μ≤K and Λ=0, according to Definition 1, we find that S1(L)=d(L,L)−d(L,L)=0, and ˜ρL(t)=(˜A(t,A0),0) is a predator-extinction periodic solution of system (2.2).
According to Lemma 2, we have
P(A,B)=rA[1−(AK)α]−τ1A2BA2+β=:Af1(A,B), |
Q(A,B)=(τ2A2A2+β−ω)B=:Bf2(A,B), |
a(A,B)=−c1hA,b(A,B)=−c2hB+Λ,ϕ(A,B)=μA+(1−μ)B. |
By direct calculations, we get
∂P∂A=f1(A,B)+Af1A(A,B),∂Q∂B=f2(A,B), |
∂a∂A=−c1h,∂b∂B=−c2h,∂ϕ∂A=μ,∂ϕ∂B=1−μ. |
Whereupon, we can obtain
Δ1=μ(1−c2h)P++(1−μ)(1−c1h)Q+μP+(1−μ)Q=(1−c1h)(1−c2h)(1−(1−c1h)IKμ)α(1−(IKμ)α)=(1−c1h)(1−c2h)(Kαμα−(1−c1h)αIα)Kαμα−Iα. |
Moreover,
∫˜T0+[∂P∂A+∂Q∂B](˜A,0)dt=∫˜T0+f1(˜A,0)+˜Af1A(˜A,0)+f2(˜A)dt=∫I/μ(1−c1)I/μ1˜AdA+∫˜T0˜Af1A(˜A,0)+f2(˜A)dt=−ln(1−c1h)+∫˜T0˜Af1A(˜A,0)+f2(˜A)dt, |
where ˜A=˜A(t,A0), and then
μ2=Δ1exp(∫˜T0[∂P∂A+∂Q∂B](˜A(t),0)dt)=(1−c1h)(1−c2h)(Kαμα−(1−c1h)αIα)Kαμα−Iαexp[−ln(1−c1h)+∫˜T0˜Af1A(˜A,0)+f2(˜A)dt]. |
Thus, when c2>˜c2, we have μ2<1 and the periodic solution ˜ρL(t)=(˜A(t,A0),0) is orbitally asymptotically stable. On the other hand, if 0<c2<˜c2, we find that μ2>1, implying the instability of the periodic solution ˜ρL(t)=(˜A(t,A0),0). The proof is completed.
Assume H is a point on the phase set with dBdA|H=−μ(1−c2h)(1−c1h)(1−μ). The trajectory of the system (2.2) starting from H intersects the isoclinic line f1(A,B)=0 and the impulse set at points H′ and H″, respectively. Define some thresholds for the system (2.2) as follows:
I∗:=μA∗+(1−μ)B∗,μ∗:=B∗K−A∗+B∗,Im:=max{I|μAH″+(1−μ)BH″≥I}. | (3.9) |
Theorem 4. System (2.2) has a positive order-1 periodic solution, if one of the conditions (i)–(iii) holds: (i)Λ=0,I≤min{I∗,μK}, 0<c2<˜c2; (ii)Λ=0,μ<μ∗,μK<I<Im; (iii)0<Λ<Λm,I≤Im, where Λm=min{c2h1−μ,c1h(1−c2h)Λ(1−μ)(1−c1h)}.
Proof. (i) Referring to Theorem 1, the boundary periodic solution of (2.2) is unstable under this condition. For any P1∈ˉN∩U(L,δ), where δ is sufficiently small, the trajectory originating from P1 intersects the impulse set ˉM at point P−1 and then is instantly pulsed to the point P+1 in the phase set ˉN. Here, P+1 is above P1 as shown in Figure 3. Following Definition 1, we have
S1(P1)=d(P+1,L)−d(P1,L)>0. |
Assuming that the point P2(A2,B2)∈ˉN satisfies function (3.6), the trajectory of system (2.2) starting from P2(A2,B2) first intersects the impulse set ˉM at P−2(A−2,B−2) and then pulses to the phase set ˉN at P+2. There are three possible situations: (1)P+2=P2, (2)P+2=P_+2 and (3)P+2=¯P+2 (shown in Figure 3).
Firstly, for the case of P+2=P2, by Definition 1 we have S1(P2)=d(P+2,L)−d(P2,L)=0. This implies the solution starting from P2 forms an order-1 periodic solution of system (2.2). Suppose P+2=P_+2, then we obtain that S1(P2)=d(P_+2,L)−d(P2,L)<0. As we have already established that S1(P1)=d(P+1,L)−d(P1,L)>0, a point E∈¯P1P2 must exist such that S1(E)=0. Consequently, there exists a positive order-1 periodic solution starting from point E.
In the scenario P+2=¯P+2, we have S1(P2)=d(¯P+2,L)−d(P2,L)>0. Since the trajectory of system (2.2) starting from P2(A2,B2) is tangent to the phase set ˉN, the trajectory of system (2.2) starting from P+2 crosses ˉN once more at ¯P+2 before reaching ˉM at P+−2 and then pulsing back to the phase set at P++2. According to the properties of system (2.1), P+−2 is below P−1 and P++2 is below P+1, which leads to
S2(¯P+2)=d(P++2,L)−d(¯P+2,L)<0. |
In addition, due to the continuity of solution in the system (2.1) and for ε=d(P2,P+2)/2, we know that there must exist some δ<ε such that D∈U(P2,δ), d(D−,P−2)<ε, and d(D+,¯P+2)≤max{(1−τ1),(1−τ2)}d(D−,P−2)<0.
Consequently,
S2(D)=d(D+,L)−d(D,L)=d(¯P+2,L)−d(¯P+2,D+)−(d(P2,L)+d(D,P2))=d(¯P+2,P2)−d(¯P+2,D+)+d(D,P2)>0. |
Thus, there must exist a point E∈¯P2¯P+2 such that S2(E)=0, and the trajectory from E forms a positive order-1 periodic solution.
(ii) For this case, we can easily have that the positive equilibrium (A∗,B∗) is above the pulse line ˉM. Let the intersections of ˉM with f1(A,B)=0 and A-axis be the points Z and Z1, respectively, where point Z1 is on the right of point (K,0). Therefore, we can obtain
μ<B∗K−A∗+B∗:=μ∗. |
Because μK<I<Im, the system doesn't yield an extinction periodic solution, and all orbits starting from the phase set ˉN reach the pulse set ˉM. Similar to the proof in (i), choosing P1∈ˉN with BP1=(1−τ2h)BZ, we can confirm the existence of an order-1 periodic solution of system (2.2).
(iii) Choose P1∈ˉN with d(P1,L)=Λ, and similar to the discussion in case (i), we can obtain the existence of a positive order-1 periodic solution for (2.2). The proof is completed.
In Theorem 4, we've already proven the existence of positive order-1 periodic solution of system (2.2). In the upcoming part, we aim to study the stability of the periodic solution. For simplicity, we introduce some notations in the following.
Let's assume a positive order-1 periodic solution of (2.2) as ˉρE(t)=(ˉA(t),ˉB(t)) with period ˉT1, and the orbit Γ(E)=^EE−E, where E=(ˉA0,ˉB0)∈ˉN, E−=(ˉA1,ˉB1)∈ˉM, ˉA1≜ˉA(ˉT), ˉB1≜ˉB(ˉT), ˉA0≜(1−c1h)¯A1, ˉB0≜(1−c2h)ˉB1+Λ. Furthermore, we mark ˉP0≜P(ˉA0,ˉB0), ˉP1≜P(ˉA1,ˉB1), ˉQ0≜Q(ˉA0,ˉB0), ˉQ1≜Q(ˉA1,ˉB1).
Denote
ℵ≜ln(ˉA0ˉB0ˉA1ˉB1μˉP1+(1−μ)ˉQ1(1−c2h)μˉP+0+(1−μ)(1−c1h)ˉQ+0). |
Theorem 5. The positive order-1 periodic solution ˉρE(t)=(ˉA(t),ˉB(t)) of system (2.2) is orbitally asymptotically stable if
∫ˉT0τ1ˉAˉB(ˉA2−β)(ˉA2+β)2−αr(ˉAK)αdt<ℵ. | (3.10) |
Proof. According to Lemma 2, we have
Δ1=μ(1−c2h)ˉP+0+(1−μ)(1−c1h)ˉQ+0μˉP1+(1−μ)ˉQ1 |
and
exp(∫ˉT0[∂ˉP∂ˉA+∂ˉQ∂ˉB](ˉA,ˉB))dt=ˉA(ˉT)ˉB(ˉT)ˉA(0)ˉB(0)exp(∫ˉT0[τ1ˉAˉB(ˉA2−β)(ˉA2+β)2−αr(ˉAK)α]dt). |
When condition (3.10) is satisfied, we can obtain
|μ2|=Δ1exp(∫ˉT0[∂ˉP∂ˉA+∂ˉQ∂ˉB](ˉA,ˉB))dt<1, |
which implies that the positive order-1 periodic solution of system (2.2) is orbitally asymptotically stable.
The existence of a positive periodic solution for the system (2.2), as shown in Theorem 4, justifies the possibility of implementing a periodic harvesting strategy with a period of ˉT. Let's consider Ac(A2<Ac<A−2) and Bc:=ϱAc to represent the target densities of prey and predator, respectively. As h relies on Ac, and both h and Λ depend on μ and Ac, we assume that
Λ(μ,Ac)≜μ[Λ1+(Ac−A2)Λ2/(A−2−A2)], |
I(μ,Ac)≜μAc+(1−μ)Bc, and h≜h(Ac)=h1+(Ac−A2)h2/(A−2−A2). |
We can also infer that ˉT depends on μ and Ac. Let n1 and n2 denote the sale prices of unit prey and predator, respectively, while n3 signifies the cost per unit capture strength and n4 represents the cost of feeding predators per unit.
To assure the sustainable renewal of resources while maximizing commercial profits, it is essential to optimize the harvesting process. The profit from harvesting can be formulated as
Wprofits(μ,Ac)=n1τ1hˉA1+n2τ2hˉB1−n3h−n4Λ(μ,Ac). |
The primary objective here is to optimize the profit cycle, i.e.,
maxWprofits(μ,Ac)ˉT(μ,Ac), | (3.11) |
where A2<Ac<A−2 and 0≤μ≤1.
By solving the optimization problem (3.11), we can deduce the optimal harvest level A∗c and weight μ∗, as well as the optimal predator release amount Λ∗, the optimal harvesting intensity h∗ and the optimal harvesting period ˉT∗. Note that these results depend on ni, where i=1,2,3,4.
This section presents some numerical simulations to validate the conclusions drawn in section 3. We adopt the following model parameters for numerical simulations:
r=0.25,τ1=0.33,τ2=0.2,α=0.5,ω=0.1,β=3.5,K=10. |
From (3.1) we can directly compute the positive equilibrium P∗(A∗,B∗)≈(1.87,1.61) with τ2=0.2,ω=0.1,β=3.5, τ2K2K2+β≈0.193>ω. Under these conditions, the positive equilibrium point P∗ of system (2.1) is globally asymptotically stable, as illustrated in Figure 4(a)–(c). This observation aligns with the conclusion in Theorem 2. However, if ω=0.3, then τ2K2K2+β≈0.193<ω, the positive equilibrium point disappears, and the boundary equilibrium point Pk=(K,0) becomes globally asymptotically stable, as shown in Figure 4(d)–(f). This finding is consistent with Theorem 1.
For the complex pulse system (2.2) we selected the following harvesting parameters:
c1=0.7,c2=0.3,h=0.8. |
With Λ=0,I=1.5,μ=0.5 satisfying I≤μK=5 and c2=0.3 satisfying c2>˜c2≈0.061, a predator-extinction periodic solution exists for system (2.2). As c2>˜c2, this period solution is orbitally asymptotically stable, as depicted in Figure 5(a)–(c). This aligns with the conclusion of Theorem 3. If, however, 0<c2<˜c2 (for instance c2=0.04), the predator-extinction periodic solution becomes unstable, and a positive periodic solution emerges, as shown in Figure 5(d)–(f). This corresponds to the positive periodic solution for the case (i) in Theorem 4.
From (3.9), we derive μ∗≈0.17,Im≈1.95. Choosing μ=0.1,I=1.5 satisfies μ<μ∗ and μK<I<Im. Under these conditions, the predator-extinction periodic solution does not exist. Figure 6(a) shows the existence of a positive periodic solution, which is orbitally asymptotically stable. When I=1.5,μ=0.4>μ∗, the predator-extinction periodic solution emerges while the positive periodic solution disappears (see Figure 6(b)). It's worth mentioning that a positive period solution also exists when I=1.5,μ=0.2 (see Figure 6(c)), suggesting that the conditions in Theorem 4(ii) are necessary but not sufficient. If I=1.98>Im,μ=0.1, the system's periodic behavior ceases, and the trajectory approaches the positive equilibrium point P∗(1.87,1.61) closely (see Figure 6(d)).
For Λ≠0, with fixed parameters I=1.5<Im,μ=0.5 and choosing Λ=0.8<Λm, the system (2.2) exhibits a positive order-1 periodic solution (see Figure 7(a), Theorem 4 (iii)). If Λ = 1.32>Λm while keeping other parameters unchanged, the predator population surpasses the threshold and the periodic solution becomes unstable, causing the prey population to become extinct after a finite number of pulses (Figure 7(c) and (e)). In order to maintain symbiosis between two populations, we can decrease prey harvesting and predator release while increasing the capture of adult predators. Setting c1=0.3,Λ=1,c2=0.6 results in the disappearance of the positive periodic solution for system (2.2) and the trajectory converges to the positive equilibrium point P∗(1.87,1.61) (see Figure 7 (b)).
When the release amount of predator cubs is reduced to Λ=1.3 (which is less than or equal to Λm) and the effective capture rate of the prey is set to c1=0.4, system (2.2) exhibits a bistable phenomenon, as seen in Figure 8(a) and (b). This system features both a locally asymptotically stable equilibrium point and a positive periodic solution. This implies that in the system (2.2), a pulse effect will not manifest if the weighted sum of the two populations is insufficient to activate it. However, a positive equilibrium does exist.
This study presents and examines a Gilpin-Ayala predator-prey system with state feedback weighted harvest strategy and a Holling Ⅲ type functional response. Through the construction of two Lyapunov functions, we have primarily studied the global stability of both the coexistence equilibrium point P∗ and the boundary equilibrium point Pk. The position of the positive equilibrium point is influenced by the magnitude of the exponential factor α, while its stability is predominantly affected by τ2, ω and β (refer to Theorem 1, Figure 4). Aiming to optimize harvest economic value and support sustainable ecological development, we introduced a weighted harvest strategy and initiating harvesting once the weighted sum of the two populations reached a predetermined threshold, causing complex dynamic effects on the predator-prey model.
For the complex pulse harvesting system, our focus lay on analyzing the existence and stability of predator-extinction periodic solutions and positive order-1 periodic solutions. In the absence of predator cub release (Λ=0), a predator-extinction periodic solution occurs when I<μK and B(0)=0. Further, this solution is asymptotically stable if c2>˜c2, as demonstrated in Theorem 3 (see Figure 5); that is, when no predator seedlings are placed and harvesting parameter c2 meets the threshold condition, the predator population will become extinct. From the perspective of ecological balance and commercial value, we do not want any species to become extinct. To this end, options include introducing predator pups into the environment or reducing the predator capture rate c2. Initially, without releasing predator cubs (Λ=0) and keeping c2<˜c2, the harvesting system (2.2) yields an asymptotically stable positive order-1 periodic solution (refer to Theorem 4, Figure 5). Moreover, when predator cubs are released, a positive order-1 periodic solution exists provided I≤Im and Λ<Λm (see Theorem 4, Figures 6 and 7). It indicates that, due to human harvesting action, the number of individuals in the area will decrease, which is not conducive to population growth and harvesting. Proper supplementation of predator seedlings can help improve population development and increase economic benefits.
A clear demonstration of bistability in system (2.2) is evident from Figures 8. This crucial observation underlines the system's sensitivity to initial conditions. In the future, we hope to prove it theoretically through mathematical methods. Due to the complexity of the switching system, only the local stability of the positive order-1 periodic solution was proven in Theorem 5, but its global stability cannot be determined using our current methods. In the future, we will try appropriate methods to study it, such as constructing Lyapunov function methods. In addition, due to a more comprehensive exploration of the predator-prey relationship in nature, we will incorporate the Allee effect, time delay effect and optimal control into the model in future work.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the Program for Science & Technology Innovation Talents in Universities of Henan Province (No. 21HASTIT026) and the Scientific Research Foundation of Graduate School of Xinyang Normal University (No. 2022KYJJ055).
The authors declare no conflict of interest regarding the publication of this paper.
[1] |
F. Ayala, M. Gilpin, J. Ehrenfeld, Competition between species: theoretical models and experimental tests, Theor. Popul. Biol., 4 (1973), 331–356. https://doi.org/10.1016/0040-5809(73)90014-2 doi: 10.1016/0040-5809(73)90014-2
![]() |
[2] |
M. Amdouni, J. Alzabut, M. Samei, W. Sudsutad, C. Thaiprayoon, A generalized approach of the Gilpin-Ayala model with fractional derivatives under numerical simulation, Mathematics, 10 (2022), 3655. https://doi.org/10.3390/MATH10193655 doi: 10.3390/MATH10193655
![]() |
[3] |
S. Ai, J. Li, J. Yu, B. Zheng, Stage-structured models for interactive wild and periodically and impulsively released sterile mosquitoes, Discrete Cont. Dyn.-B, 27 (2022), 3039–3052. https://doi.org/10.3934/dcdsb.2021172 doi: 10.3934/dcdsb.2021172
![]() |
[4] |
I. Boubekri, H. Mazurek, A. Djebar, R. Amara, Harnessing Fishers' local knowledge and their perceptions: opportunities to improve management of coastal fishing in Mediterranean marine protected areas, J. Environ. Manage., 344 (2023), 118456. https://doi.org/10.1016/J.JENVMAN.2023.118456 doi: 10.1016/J.JENVMAN.2023.118456
![]() |
[5] |
J. Chen, J. Huang, S. Ruan, J. Wang, Bifurcations of invariant tori in predator-prey models with seasonal prey harvesting, SIAM J. Appl. Math., 73 (2013), 1876–1905. https://doi.org/10.1137/120895858 doi: 10.1137/120895858
![]() |
[6] | L. Chen, Pest control and geometric theory of semi-continuous dynamical system, J. Beihua Univ., 12 (2011), 9–11. |
[7] |
M. Costa, E. Kaszkurewicz, A. Bhaya, L. Hsu, Achieving global convergence to an equilibrium population in predator-prey systems by the use of a discontinuous harvesting policy, Ecol. Model., 128 (2000), 89–99. https://doi.org/10.1016/S0304-3800(99)00220-3 doi: 10.1016/S0304-3800(99)00220-3
![]() |
[8] |
S. Debnath, P. Majumdar, S. Sarkar, U. Ghosh, Global dynamics of a prey-predator model with Holling type Ⅲ functional response in the presence of harvesting, J. Biol. Syst., 30 (2022), 225–260. https://doi.org/10.1142/S0218339022500073 doi: 10.1142/S0218339022500073
![]() |
[9] |
E. Gonzˇslez-Olivares, A. Rojas-Palma, Multiple limit cycles in a gause predator-prey model with Holling Ⅲ functional response and Allee effect on prey, Bull. Math. Biol., 73 (2011), 1378–1397. https://doi.org/10.1007/s11538-010-9577-5 doi: 10.1007/s11538-010-9577-5
![]() |
[10] |
H. Guo, X. Song, L. Chen, Qualitative analysis of a korean pine forest model with impulse thinning measure, Appl. Math. Comput., 234 (2014), 203–213. https://doi.org/10.1016/j.amc.2014.02.034 doi: 10.1016/j.amc.2014.02.034
![]() |
[11] |
C. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, The Memoirs of the Entomological Society of Canada, 97 (1965), 5–60. https://doi.org/10.4039/entm9745fv doi: 10.4039/entm9745fv
![]() |
[12] |
M. Huang, S. Liu, X. Song, L. Chen, Periodic solutions and homoclinic bifurcation of a predator-prey system with two types of harvesting, Nonlinear Dyn., 73 (2013), 815–826. https://doi.org/10.1007/s11071-013-0834-7 doi: 10.1007/s11071-013-0834-7
![]() |
[13] |
M. Huang, X. Yu, S. Liu, X. Song, Dynamical behavior of a mosquito population suppression model composed of two sub-models, Int. J. Biomath., 16 (2023), 2250126. https://doi.org/10.1142/S1793524522501261 doi: 10.1142/S1793524522501261
![]() |
[14] |
M. Huang, X. Yu, S. Liu, Modeling and analysis of release strategies of sterile mosquitoes incorporating stage and sex structure of wild ones, Electron. Res. Arch., 31 (2023), 3895–3914. https://doi.org/10.3934/era.2023198 doi: 10.3934/era.2023198
![]() |
[15] |
M. Huang, X. Yu, Dynamic analysis of a mosquito population model with stage structure and periodic releases of sterile ones, AIMS Mathematics, 8 (2023), 18546–18565. https://doi.org/10.3934/math.2023943 doi: 10.3934/math.2023943
![]() |
[16] |
J. Jiao, L. Chen, S. Cai, Dynamical analysis of a biological resource management model with impulsive releasing and harvesting, Adv. Differ. Equ., 2012 (2012), 9. https://doi.org/10.1186/1687-1847-2012-9 doi: 10.1186/1687-1847-2012-9
![]() |
[17] |
Y. Jiang, Z. Liu, J. Yang, Y. Tan, Dynamics of a stochastic Gilpin-Ayala population model with Markovian switching and impulsive perturbations, Adv. Differ. Equ., 2020 (2020), 530. https://doi.org/10.1186/s13662-020-02900-w doi: 10.1186/s13662-020-02900-w
![]() |
[18] |
G. Kaniadakis, Novel predator-prey model admitting exact analytical solution, Phys. Rev. E, 106 (2022), 044401. https://doi.org/10.1103/PHYSREVE.106.044401 doi: 10.1103/PHYSREVE.106.044401
![]() |
[19] |
V. Křivan, The Lotka-Volterra predator-prey model with foraging-predation risk trade-offs, Am. Nat., 170 (2007), 771–782. https://doi.org/10.1086/522055 doi: 10.1086/522055
![]() |
[20] |
F. Rihan, H. Alsakaji, C. Rajivganthi, Stability and Hopf bifurcation of three-species prey-predator system with delays and Allee effect, Complexity, 2020 (2020), 7306412. https://doi.org/10.1155/2020/7306412 doi: 10.1155/2020/7306412
![]() |
[21] |
F. Souna, P. Tiwari, M. Belabbas, Y. Menacer, A predator-prey system with prey social behavior and generalized Holling Ⅲ functional response: role of predator-taxis on spatial patterns, Math. Method. Appl. Sci., 46 (2023), 13991–14006. https://doi.org/10.1002/MMA.9300 doi: 10.1002/MMA.9300
![]() |
[22] |
K. Sun, T. Zhang, Y. Tian, Theoretical study and control optimization of an integrated pest management predator-prey model with power growth rate, Math. Biosci., 279 (2016), 13–26. https://doi.org/10.1016/j.mbs.2016.06.006 doi: 10.1016/j.mbs.2016.06.006
![]() |
[23] |
S. Tang, B. Tang, A. Wang, Y. Xiao, Holling Ⅱ predator-prey impulsive semi-dynamic model with complex Poincar map, Nonlinear Dyn., 81 (2015), 1575–1596. https://doi.org/10.1007/s11071-015-2092-3 doi: 10.1007/s11071-015-2092-3
![]() |
[24] |
Y. Tian, Y. Gao, K. Sun, Global dynamics analysis of instantaneous harvest fishery model guided by weighted escapement strategy, Chaos Soliton. Fract., 164 (2022), 112597. https://doi.org/10.1016/j.chaos.2022.112597 doi: 10.1016/j.chaos.2022.112597
![]() |
[25] |
Y. Tian, Y. Gao, K. Sun, Qualitative analysis of exponential power rate fishery model and complex dynamics guided by a discontinuous weighted fishing strategy, Commun. Nonlinear Sci., 118 (2023), 107011. https://doi.org/10.1016/j.cnsns.2022.107011 doi: 10.1016/j.cnsns.2022.107011
![]() |
[26] |
M. Vasilova, Asymptotic behavior of a stochastic Gilpin-Ayala predator-prey system with time-dependent delay, Math. Comput. Model., 57 (2013), 764–781. https://doi.org/10.1016/j.mcm.2012.09.002 doi: 10.1016/j.mcm.2012.09.002
![]() |
[27] |
M. Vasilova, M. Jovanvic, Stochastic Gilpin-Ayala competition model with infinite delay, Appl. Math. Comput., 217 (2011), 4944–4959. https://doi.org/10.1016/j.amc.2010.11.043 doi: 10.1016/j.amc.2010.11.043
![]() |
[28] |
V. Volterra, Fluctuations in the abundance of a species considered mathematically, Nature, 118 (1926), 558–560. https://doi.org/10.1038/118558a0 doi: 10.1038/118558a0
![]() |
[29] |
W. Wei, W. Xu, J. Liu, Y. Song, S. Zhang, Stochastic bifurcation and Break-out of dynamic balance of predator-prey system with Markov switching, Appl. Math. Model., 117 (2023), 563–576. https://doi.org/10.1016/J.APM.2022.12.034 doi: 10.1016/J.APM.2022.12.034
![]() |
[30] |
H. Xu, T. Zhang, H. Cheng, Nonlinear control ecological model with complex discrete map, Commun. Nonlinear Sci., 118 (2023), 107019. https://doi.org/10.1016/j.cnsns.2022.107019 doi: 10.1016/j.cnsns.2022.107019
![]() |
[31] |
J. Xu, M. Huang, X. Song, Dynamics analysis of a two-species competitive system with state feedback impulsive control, Int. J. Biomath., 13 (2020), 2050007. https://doi.org/10.1142/S1793524520500072 doi: 10.1142/S1793524520500072
![]() |
[32] |
M. Zhang, L. Chen, Z. Li, Homoclinic bifurcation of a state feedback impulsive controlled prey-predator system with Holling-Ⅱ response, Nonlinear Dyn., 98 (2019), 929–942. https://doi.org/10.1007/s11071-019-05235-8 doi: 10.1007/s11071-019-05235-8
![]() |
[33] |
Q. Zhang, S. Tang, X. Zou, Rich dynamics of a predator-prey system with state-dependent impulsive controls switching between two means, J. Differ. Equations, 364 (2023), 336–377. https://doi.org/10.1016/J.JDE.2023.03.030 doi: 10.1016/J.JDE.2023.03.030
![]() |
[34] |
X. Zeng, L. Liu, W. Xie, Existence and uniqueness of the positive steady state solution for a Lotka-Volterra predator-prey model with a crowding term, Acta. Math. Sci., 40 (2020), 1961–1980. https://doi.org/10.1007/s10473-020-0622-7 doi: 10.1007/s10473-020-0622-7
![]() |