
In this paper, a prey-predator model with modified Leslie-Gower and simplified Holling-type Ⅳ functional responses is proposed to study the dynamic behaviors. For the deterministic system, we analyze the permanence of the system and the stability of the positive equilibrium point. For the stochastic system, we not only prove the existence and uniqueness of global positive solution, but also discuss the persistence in mean and extinction of the populations. In addition, we find that stochastic system has an ergodic stationary distribution under some parameter constraints. Finally, our theoretical results are verified by numerical simulations.
Citation: Lin Li, Wencai Zhao. Deterministic and stochastic dynamics of a modified Leslie-Gower prey-predator system with simplified Holling-type Ⅳ scheme[J]. Mathematical Biosciences and Engineering, 2021, 18(3): 2813-2831. doi: 10.3934/mbe.2021143
[1] | Yongxin Gao, Shuyuan Yao . Persistence and extinction of a modified Leslie-Gower Holling-type Ⅱ predator-prey stochastic model in polluted environments with impulsive toxicant input. Mathematical Biosciences and Engineering, 2021, 18(4): 4894-4918. doi: 10.3934/mbe.2021249 |
[2] | Yan Zhang, Shujing Gao, Shihua Chen . Modelling and analysis of a stochastic nonautonomous predator-prey model with impulsive effects and nonlinear functional response. Mathematical Biosciences and Engineering, 2021, 18(2): 1485-1512. doi: 10.3934/mbe.2021077 |
[3] | Fathalla A. Rihan, Hebatallah J. Alsakaji . Analysis of a stochastic HBV infection model with delayed immune response. Mathematical Biosciences and Engineering, 2021, 18(5): 5194-5220. doi: 10.3934/mbe.2021264 |
[4] | Xueqing He, Ming Liu, Xiaofeng Xu . Analysis of stochastic disease including predator-prey model with fear factor and Lévy jump. Mathematical Biosciences and Engineering, 2023, 20(2): 1750-1773. doi: 10.3934/mbe.2023080 |
[5] | Lazarus Kalvein Beay, Agus Suryanto, Isnani Darti, Trisilowati . Hopf bifurcation and stability analysis of the Rosenzweig-MacArthur predator-prey model with stage-structure in prey. Mathematical Biosciences and Engineering, 2020, 17(4): 4080-4097. doi: 10.3934/mbe.2020226 |
[6] | Yansong Pei, Bing Liu, Haokun Qi . Extinction and stationary distribution of stochastic predator-prey model with group defense behavior. Mathematical Biosciences and Engineering, 2022, 19(12): 13062-13078. doi: 10.3934/mbe.2022610 |
[7] | Mengyun Xing, Mengxin He, Zhong Li . Dynamics of a modified Leslie-Gower predator-prey model with double Allee effects. Mathematical Biosciences and Engineering, 2024, 21(1): 792-831. doi: 10.3934/mbe.2024034 |
[8] | Claudio Arancibia–Ibarra, José Flores . Modelling and analysis of a modified May-Holling-Tanner predator-prey model with Allee effect in the prey and an alternative food source for the predator. Mathematical Biosciences and Engineering, 2020, 17(6): 8052-8073. doi: 10.3934/mbe.2020408 |
[9] | Eduardo Gonzalez-Olivares, Alejandro Rojas-Palma . Global stability in a modified Leslie-Gower type predation model assuming mutual interference among generalist predators. Mathematical Biosciences and Engineering, 2020, 17(6): 7708-7731. doi: 10.3934/mbe.2020392 |
[10] | Manoj K. Singh, Brajesh K. Singh, Poonam, Carlo Cattani . Under nonlinear prey-harvesting, effect of strong Allee effect on the dynamics of a modified Leslie-Gower predator-prey model. Mathematical Biosciences and Engineering, 2023, 20(6): 9625-9644. doi: 10.3934/mbe.2023422 |
In this paper, a prey-predator model with modified Leslie-Gower and simplified Holling-type Ⅳ functional responses is proposed to study the dynamic behaviors. For the deterministic system, we analyze the permanence of the system and the stability of the positive equilibrium point. For the stochastic system, we not only prove the existence and uniqueness of global positive solution, but also discuss the persistence in mean and extinction of the populations. In addition, we find that stochastic system has an ergodic stationary distribution under some parameter constraints. Finally, our theoretical results are verified by numerical simulations.
In the 1960s, mite pests broke out on fruit trees in Washington state of the United States, resulting in a large area of production reduction. Mite pests can not only cause the leaves of fruit trees to turn green and fall off, but also cause the trees to weaken, affecting the flowering and fruiting of the next year. Mite pests have a high reproductive rate, one female adult mite can reproduce more than 10,000 larvae within a month under suitable conditions. At the same time, mite pests are easy to develop resistance, so it is difficult to control them. Washington state is a major quality apple producing area in the United States, and its cultivation technology and pest control level are in the leading position in the United States. In 1969, Dr. S. C. Hoyt proposed a comprehensive pest management strategy combining chemical control with biological control[1]. The main pest of fruit trees, Tetranychus mcdanieli was controlled by protecting and utilizing its natural enemy, Typhlodromus occidentalis Nesbitt. In 1988, D. J. Wollkind et al.[2] studied a mite predator-prey model with Leslie term and Holling type function, and discussed the dependence between the sub-stability of the system and temperature. The form of the model is as follows:
{˙x(t)=ax(t)(1−x(t)K)−kx(t)y(t)r1+x(t),˙y(t)=cy(t)(1−y(t)r2x(t)), | (1.1) |
where x(0)=x0,y(0)=y0. x(t) and y(t) stand for the population density of prey and predator at time t, respectively. a and c are the corresponding intrinsic growth rates of the two populations. k is the maximum per capita consumption rate of prey caused by the predator, and r1 denotes the density of prey required to reach half of this rate. K and r2 represent the environmental carrying capacity of the prey and the conversion factor of prey into predator, respectively. a, c, k, r1, r2, K>0. The model (1.1) adopts Holling-type II functional response, which is a common response of predatory mites. In 1997, J.B. Collings[3] studied the effects of different functional responses on the dynamic behavior of mite predator-prey interaction models. For Holling-type I, II, and III functional responses, the models have similar dynamic properties in a given temperature range, including bifurcation and stability. If Holling-type IV is used to replace Holling-type II functional response in the system (1.1), the model still has similar properties when the interaction level of prey interference is low. However, when the level of interference is higher, there will be significant differences, such as the emergence of bistability and attractors. Holling-type IV functional response[4] is the following rational nonmonotonic function:
h(x)=kxax2+bx+c, |
also known as Monod-Haldane function. The parameter a measures the level of prey interference with predation. The function reflects that the unit predation rate of predator reaches the maximum with the increase of prey density, and then decreases when it exceeds the critical value. Holling-type IV functional response is mainly used to describe the inhibition effect of microbial population, the group defense of prey population, and the agammaegation effect of biological population, such as a fish school that is several miles long may make it difficult for predators to attack.
It is pointed out in [5] that when the density of apple tree pests (such as T. mcdanieli) is high, they will produce a kind of webbing which is known to interfere with predators by decreasing their walking speed and reducing their chances of contacting the prey, so as to effectively reduce the probability of being preyed upon. Sometimes predatory mites even starve in front of the prey. It is reasonable to choose Holling-type IV functional response if we assume that the prey mite and webbing density are directly related. This paper discusses the population dynamics of prey mite interference with predation, which is of great significance to the control of fruit pests.
In 1981, W. Sokol et al.[6] adopted the simplified Holling-type Ⅳ functional response function h(x)=kxr1+x2 to model and claimed that it was better consistent with the experimental data. Similarly, [7,8] used this response function to get the following mites predator-prey model:
{˙x(t)=ax(t)(1−x(t)K)−kx(t)y(t)r1+x2(t),˙y(t)=cy(t)(1−y(t)r2x(t)). | (1.2) |
In the second equation of model (1.2), y(t)r2x(t) is called the Leslie-Gower term[9,10], which reflects the inverse relationship between the decrease of the predator's number and the per capita availability of their favorite prey. However, due to the influence of environment and season on population growth and reproduction, the number of prey may decrease sharply in a certain period of time. In order to survive, the predator will catch and feed on other prey to get enough food for growth. Considering that the number of prey is insufficient, reference [11] {has} obtained a modified Leslie-Gower term by adding a positive number to the denominator of the Leslie-Gower term: y(t)r2+x(t), where r2 expresses the quantity of available food different from the predator's favorite prey. By coupling the modified Leslie-Gower term with the system (1.2) to get the mites model as follows[12]:
{˙x(t)=x(t)(a−bx(t)−ky(t)r1+x2(t)),˙y(t)=y(t)(c−py(t)r2+x(t)), | (1.3) |
where b and p signify the competition strength in the prey individuals and the maximum per capita reduction rate of the predator, respectively. a,b,c,k,r1,r2,p>0.
Nevertheless, biological populations are inevitably affected by many environmental factors, such as temperature, sunlight, humidity and so on[13]. These disturbances are random, little and independent, and white noise can be used to express the influence of these disturbances. Many scholars have studied the dynamic behavior of stochastic system by introducing environmental white noise into deterministic model[14,15,16,17,18,19,20,21]. This paper assumes that the intrinsic growth rates a and c of system (1.3) are disturbed by environmental white noise: a→a+α˙B1(t), c→c+β˙B2(t). The following stochastic system is obtained:
{dx(t)=x(t)(a−bx(t)−ky(t)r1+x2(t))dt+αx(t)dB1(t),dy(t)=y(t)(c−py(t)r2+x(t))dt+βy(t)dB2(t), | (1.4) |
where B1(t) and B2(t) are independent Brownian motion, α2 and β2 represent the intensity of white noise.
In this paper, we will compare the deterministic system (1.3) with the stochastic system (1.4), focusing on the long-time dynamics of the populations. In Section 2, we first discuss the permanence of deterministic system (1.3), and then obtain the stability conditions of positive equilibrium by using Routh-Hurwitz criterion and constructing an appropriate Lyapunov function. For the stochastic system (1.4), we demonstrate the existence and uniqueness of the global positive solution. Then the conditions for the persistence in mean and extinction of populations are discussed. And the parameter constraint criterion for the existence of the ergodic stationary distribution is found, which will be presented in Section 3. In Section 4, numerical simulations are used to verify our theoretical results.
Definition 2.1. [22] If there exists 0<M<N such that for any solution (x(t),y(t))∈R2+ of the system (1.3), there is
M≤min{limt→+∞infx(t),limt→+∞infy(t)}≤max{limt→+∞supx(t),limt→+∞supy(t)}≤N, |
then the system (1.3) is permanent.
Theorem 2.1. Let (x(t),y(t))∈R2+ be the solution of the system (1.3) satisfying the initial condition (x(0),y(0))∈R2+, then
1b[a−ck(r2b+a)pr1b]≤limt→+∞infx(t)≤limt→+∞supx(t)≤ab, |
cr2p≤limt→+∞infy(t)≤limt→+∞supy(t)≤cp(r2+ab). |
The system (1.3) is permanent if apr1b>ck(r2b+a).
Proof. Derived from system (1.3)
dxdt≤x(a−bx),dydt≥y(c−pyr2). |
On the basis of Lemma 2.3 in [23], we have
limt→+∞supx(t)≤ab, | (2.1) |
limt→+∞infy(t)≥cr2p. | (2.2) |
Therefore, for any ε1>0, there exists T1>0; when t≥T1, there is x≤ab+ε1. So when t≥T1, we can get
dydt≤y(c−pyr2+ab+ε1). |
In accordance with Lemma 2.3 in [23]
limt→+∞supy(t)≤c(r2+ab+ε1)p, |
when ε1→0, we obtain that
limt→+∞supy(t)≤c(r2+ab)p. | (2.3) |
From (2.3), for any ε2, there exists T2>0; when t≥T2, there is y≤c(r2+ab)p+ε2. Hence, when t≥T2
dxdt≥x(a−bx−kyr1)≥x[a−bx−k(cr2b+ca+bpε2)pr1b]. |
According to Lemma 2.3 in [23]
limt→+∞infx(t)≥1b[a−k(cr2b+ca+pbε2)pr1b], |
when ε2→0, we can see that
limt→+∞infx(t)≥1b[a−ck(r2b+a)pr1b]. | (2.4) |
Consequently, we have known that the system (1.3) is permanent if apr1b>ck(r2b+a).
In this section, we analyze the types of equilibrium points of the system(1.3). E1=(0,0) is the trivial equilibrium point. E2=(ab,0) and E3=(0,cr2p) are two boundary equilibrium points. The internal equilibrium point E∗=(x∗,y∗) is the intersection point of two isoclinic lines kyr1+x2=a−bx and y=c(r2+x)p in the interior of the first quadrant. According to the equations
{y=1k(a−bx)(x2+r1),y=cp(x+r2), |
we can obtain
H(x)=bx3−ax2+(br1+kcp)x−(ar1−kcr2p)=0. | (2.5) |
The positive equilibrium points of the system (1.3) depend on the existence of the positive roots of equation (2.5). According to Ref. [12], it is concluded that:
(i) If D=ar1−kcr2p>0, equation (2.5) may have one, two or three positive roots in different parameter ranges.
(ii) If D=ar1−kcr2p=0, equation (2.5) may have one or two positive roots or no positive real roots in different parameter ranges.
(iii) If D=ar1−kcr2p<0, equation (2.5) may have one or two positive roots or no positive real roots in different parameter ranges.
Therefore, the model (1.3) can have at most three positive equilibrium points, which may be saddle point, node, focus or center under different conditions. By means of numerical simulations, the authors revealed that the system had abundant dynamic properties including noninfinitesimal limit cycle and bifurcation (for details, please refer to Ref. [12]).
In conclusion, if D=ar1−kcr2p>0, the model (1.3) has at least one positive equilibrium point. Next, we focus on the stability of the positive equilibrium point E∗=(x∗,y∗).
For the convenience of discussion, the following assumptions are given:
(H1): a>ckr2pr1;
(H2): a<ckr2pr1;
(H3): 2kx∗y∗(r1+x∗2)2<min{b+cx∗,b+ckp(r1+x∗2)};
(H4): −b+ky∗(x∗+ab)r1(r1+x∗2)<−y∗(r2+ab)4p(r2+x∗)(p2r22+k2r12).
Based on the analysis of the types of equilibrium points in [12], we can know that: E1=(0,0) is an unstable node; E2=(ab,0) is a saddle point; E3=(0,cr2p) is a saddle point if the condition (H1) holds; E3=(0,cr2p) is a stable node if the condition (H2) holds. We begin to analyze the stability of the positive equilibrium point E∗=(x∗,y∗) in the following content.
Theorem 2.2. If conditions (H1) and (H3) hold, thus E∗=(x∗,y∗) is locally asymptotically stable. Furthermore, if condition (H4) holds, then E∗=(x∗,y∗) is globally asymptotically stable.
Proof. The Jacobian matrix at E∗=(x∗,y∗) is
J(x∗,y∗)=(−bx∗+2kx∗2y∗(r1+x∗2)2−kx∗r1+x∗2c2p−c). |
The characteristic equation of J(x∗,y∗) is
λ2+(c+bx∗−2kx∗2y∗(r1+x∗2)2)λ+(bx∗−2kx∗2y∗(r1+x∗2)2)c+c2p⋅kx∗r1+x∗2=0. |
By (H3) we can obtain that c+bx∗−2kx∗2y∗(r1+x∗2)2>0, (bx∗−2kx∗2y∗(r1+x∗2)2)+cp⋅kx∗r1+x∗2>0. Thus, according to the Routh-Hurwitz criterion, E∗=(x∗,y∗) is locally asymptotically stable.
Next, we study the global asymptotic stability of E∗, as long as it is globally attractive. Define
V(x,y)=∫xx∗ξ−x∗ξdξ+h∫yy∗η−y∗ηdη, |
where h is a positive number, as defined in the following text. We can see that V(x,y) is a non-negative function, V(x,y)=0 if and only if (x,y)=(x∗,y∗).
dVdt=x−x∗x⋅x(a−bx−kyr1+x2)+h⋅y−y∗y⋅y(c−pyr2+x)=(x−x∗)(a−bx−kyr1+x2−a+bx∗+ky∗r1+x∗2)+h(y−y∗)(c−pyr2+x−c+py∗r2+x∗)=−b(x−x∗)2+k(x−x∗)y∗(x+x∗)(x−x∗)−(r1+x∗2)(y−y∗)(r1+x∗2)(r1+x2)+ph(y−y∗)y∗(x−x∗)−(r2+x∗)(y−y∗)(r2+x∗)(r2+x). |
Choose h=r2+x∗y∗, hence
dVdt=(x−x∗)2[−b+ky∗(x+x∗)(r1+x∗2)(r1+x2)]+(x−x∗)(y−y∗)(pr2+x−kr1+x2)−p(r2+x∗)(y−y∗)2y∗(r2+x). |
The above equation can be written as
dVdt=−(x−x∗,y−y∗)(−A−B−B C)(x−x∗y−y∗), |
where
A=−b+ky∗(x+x∗)(r1+x∗2)(r1+x2), |
B=12(pr2+x−kr1+x2), |
C=p(r2+x∗)y∗(r2+x). |
Then dVdt<0 if the above matrix is a positive definite matrix(see Ref.[11]). According to the judgment of positive definite matrix, we only need to prove that its leading principal minors are positive. That is to say:
(i):−A>0,
(ii):−AC−B2>0.
In fact, if (H4) holds, then A<0 is obviously true.
And
AC+B2=[−b+ky∗(x+x∗)(r1+x∗2)(r1+x2)]⋅p(r2+x∗)y∗(r2+x)+14(pr2+x−kr1+x2)2<[−b+ky∗(ab+x∗)r1(r1+x∗2)]⋅p(r2+x∗)y∗(r2+ab)+14(p2r22+k2r12). |
From the condition (H4), AC+B2<0 clearly. Hence, dVdt<0 for any point except the equilibrium point E∗=(x∗,y∗), and dVdt|(x∗,y∗)=0. In the end, we can conclude that E∗ is globally asymptotically stable.
The global stability of the positive equilibrium point in the deterministic system means that under certain conditions, the two populations will coexist with each other over time and the population density will be stable near the positive equilibrium point.
In this section, we introduce environmental white noise into the deterministic system (1.3) to obtain the stochastic system (1.4). Next, we consider the influence of white noise on the dynamic behavior of the system (1.4).
We consider whether system (1.4) has a unique global positive solution for any given initial value (x(0),y(0))∈R2+. Generally, we will first prove that there exists a unique local positive solution, and then show that the solution is also global.
Theorem 3.1. For any given initial value (x(0),y(0))∈R2+, the stochastic system (1.4) has unique global positive solution (x(t),y(t))∈R2+ for all t≥0 and exists in R2+ with probability 1.
Proof. The equation coefficients of the stochastic system (1.4) neither satisfy the linear growth condition, nor the local Lipschitz condition[24]. Let u(t)=lnx(t),v(t)=lny(t), then the system (1.4) is transformed into the following equivalent system:
{du(t)=(a−α22−beu(t)−kev(t)r1+e2u(t))dt+αdB1(t),dv(t)=(c−β22−pev(t)r2+eu(t))dt+βdB2(t). | (3.1) |
Obviously, the coefficients of equation (3.1) satisfy the local Lipschitz condition, then (3.1) exists a unique local solution (u(t),v(t)),t∈[0,τe). τe is the explosion time which means that limt→τe|u(t)|=∞ or limt→τe|v(t)|=∞. Since the (3.1) is equivalent to the (1.4), system (1.4) has a unique local solution. On the basis of Itˆo's formula, it can be known that (x(t),y(t))=(eu(t),ev(t)) is the only local solution satisfying the initial value (x(0),y(0))∈R2+ for the stochastic system (1.4). In order to prove the solution is global, we only need to prove τe=+∞.
The following proof process is similar to [25], we can refer to the detailed proof of it. In the proof process, the most momentous thing is to construct a suitable Lyapunov function.
We define a C2-function V(x,y):R2+→¯R+,
V(x,y)=x−logx+y−logy. |
Its nonnegativity can be determined by z−logz≥1 for any z>0. In the light of Itˆo's formula to get
LV(x,y)=(x−1)(a−bx−kyr1+x2)+(y−1)(c−pyr2+x)+α22+β22≤(a+b)x+(c+kr1+pr2)y+(−a−c+α22+β22). |
Define two positive constants
h1=(a+b), h2=(c+kr1+pr2). |
Then
LV(x,y)≤h1x+h2y+(−a−c+α22+β22). |
By the inequality x≤2(x−1−logx)+log4, y≤2(y−1−logy)+log4,
LV(x,y)≤2h1(x−1−logx)+2h2(y−1−logy)+h1log4+h2log4+(−a−c+α22+β22)≤2(h1+h2)V(x,y)+(h1+h2)log4+(−a−c+α22+β22)≤h3(1+V(x,y)), | (3.2) |
where h3=max{2(h1+h2),(h1+h2)log4+(−a−c+α22+β22)}. From (3.2), we have
dV(x,y)=LV(x,y)dt+α(x−1)dB1(t)+β(y−1)dB2(t)≤h3(1+V(x,y))dt+α(x−1)dB1(t)+β(y−1)dB2(t). |
Integrating both sides from 0 to τk∧T (τk is a stopping time and T>0 is a constant, as defined in Ref. [25]) and taking the expectation
EV(x(τk∧T),y(τk∧T))≤V(x(0),y(0))+E∫τk∧T0h3(1+V(x(t),y(t)))dt≤V(x(0),y(0))+h3T+h3E∫τk∧T0V(x(t),y(t))dt≤V(x(0),y(0))+h3T+h3∫T0EV(x(τk∧t),y(τk∧t))dt. |
By Gronwall inequality
E(V(x(τk∧T),y(τk∧T)))≤(V(x(0),y(0))+h3T)eh3T. |
The following proof is the same as [25], so we omit it here.
In this section, we demonstrate that population extinction and persistence in mean will occur under some parameter constraints.
First of all, we give condition (H5): k2+4p(r2k−r1p)≤0, then there is:
Theorem 3.2. For the stochastic system (1.4), the following conclusions hold:
(i) If a<α22, c<β22, both the predator and the prey are extinct.
(ii) If a>α22, c<β22, the prey x(t) is persistent in mean and the predator y(t) is extinct.
(iii) If a<α22, c>β22, the prey x(t) is extinct and the predator y(t) is persistent in mean.
(iv) If a−α22>c−β22>0 and the condition (H5) is satisfied, both the prey x(t) and the predator y(t) are persistent in mean.
Proof. According to Itˆo's formula
dlogx(t)=(a−bx(t)−ky(t)r1+x2(t)−α22)dt+αdB1(t), |
dlogy(t)=(c−py(t)r2+x(t)−β22)dt+βdB2(t). |
Integrating both sides from 0 to t and dividing t, respectively
logx(t)t=logx(0)t+(a−α22)−bt∫t0x(s)ds−1t∫t0ky(s)r1+x2(s)ds+αt∫t0dB1(s), | (3.3) |
logy(t)t=logy(0)t+(c−β22)−1t∫t0py(s)r2+x(s)ds+βt∫t0dB2(s). | (3.4) |
(i) Applying the strong law of large number for local martingales, it can be seen from (3.3) and (3.4)
limt→+∞logx(t)t≤a−α22<0, |
limt→+∞logy(t)t≤c−β22<0. |
Consequently
limt→+∞x(t)=0, limt→+∞y(t)=0. |
Then, the prey x(t) and the predator y(t) are both extinct.
(ⅱ) From (3.4), we have limt→+∞y(t)=0. In this case, the limit system of (1.4) is
dx(t)=x(t)(a−bx(t))dt+αx(t)dB1(t). |
Therefore, on the basis of Lemma A.1 in [26]
limt→+∞1t∫t0x(s)ds=a−α22b>0, |
which means that the prey x(t) is persistent in mean.
(ⅲ) According to (3.3), we can obviously get limt→+∞x(t)=0. Moreover, the limit system of (1.4) is
dy(t)=y(t)(c−pr2y(t))dt+βy(t)dB2(t). |
Similarly
limt→+∞1t∫t0y(s)ds=r2(c−β22)p>0. |
It signifies that the predator y(t) is persistent in mean.
(ⅳ) In the light of (3.4)
logy(t)t≥logy(0)t+(c−β22)−1t∫t0py(s)r2ds+βt∫t0dB2(s). |
Thereupon, in view of Lemma 2 in [27]
limt→+∞inf1t∫t0y(s)ds≥r2(c−β22)p. |
In the same way, combining (3.4) and Lemma 2 of [28], we can obtain
limt→+∞1t∫t0py(s)r2+x(s)ds=c−β22. |
According to the condition (H5), there is kr1+x2≤pr2+x, then combining (3.3) to get
logx(t)t=logx(0)t+(a−α22)−bt∫t0x(s)ds−1t∫t0ky(s)r1+x2(s)ds+αt∫t0dB1(s)≥logx(0)t+(a−α22)−bt∫t0x(s)ds−1t∫t0py(s)r2+x(s)ds+αt∫t0dB1(s). |
Similarly, we get
limt→+∞inf1t∫t0x(s)ds≥a−α22−(c−β22)b. |
As a result, both the prey x(t) and the predator y(t) are persistent in mean.
We can testify the stability of the deterministic system by analyzing the local and global stability of the equilibrium points. However, there are no equilibrium points in the stochastic system, so we can not use the equilibrium points theory to certify the stability. Therefore, we utilize the stationary distribution to depict the stochastic weak stability. And ergodicity means that the system is persistent[25]. Under some conditions, we prove that the stochastic system (1.4) has an ergodic stationary distribution, which will be given in the following.
First, we define a parameter:
λ=(a−α22)−(c−β22)−pr2, |
then
Theorem 3.3. If the parameters of the system (1.4) satisfy c−β2>0, λ>0 and condition (H5), then system (1.4) has unique ergodic stationary distribution.
Proof. For the sake of proving Theorem 3.3, we simply need to confirm the two conditions in Lemma 2.1[29]. At first, the diffusion matrix of system (1.4) is
A=(α2x200β2y2). |
We think that U is a bounded domain, and it is apparent that there exists
M=min(x,y)∈¯U⊂R2+{α2x2,β2y2}>0, |
such that
2∑i,j=1aijξiξj=α2x2ξ12+β2y2ξ22≥M∥ξ∥2 forall (x,y)∈¯U,ξ=(ξ1,ξ2)∈R2. |
Then the first condition holds, and next we verify the second condition.
Define the C2-function ¯V(x,y):R2+→R+
¯V(x,y)=x−1−logx+1y+logy. |
Obviously, ¯V(x,y) has a unique stationary point (1,1). It is easy to calculate that (1,1) is the minimum point and the minimum value is ¯V(1,1)=1. Define the Lyapunov function
˜V(x,y)=¯V(x,y)−¯V(1,1)=V1(x,y)+V2(x,y)−1, |
where V1(x,y)=x−1−logx, V2(x,y)=1y+logy.
Utilizing Itˆo's formula and condition (H5), we have
LV1(x,y)=(x−1)(a−bx−kyr1+x2)+12α2=−bx2+(a+b)x−kxyr1+x2+kyr1+x2−(a−12α2)≤−bx2+(a+b)x−kxyr1+x2+pyr2+x−(a−12α2). |
LV2(x,y)=(−1y2+1y)⋅y(c−pyr2+x)+12β2y2(2y3−1y2)=−1y(c−β2)+pr2+x−pyr2+x+(c−12β2)≤−1y(c−β2)−pyr2+x+(c−12β2)+pr2. |
Then
L˜V(x,y)≤−bx2+(a+b)x−kxyr1+x2−1y(c−β2)−λ. |
We choose sufficiently small ε1,ε2 such that
0<ε1<λ2(a+b), | (3.5) |
0<ε2<4b(c−β2)4b+(a+b)2, | (3.6) |
0<ε12<b22b+(a+b)2, | (3.7) |
0<ε2<kε21(r1ε21+1)(a+b). | (3.8) |
And we consider bounded open set
Uε1,2={(x,y)∈R2+|ε1<x<1ε1,ε2<y<1ε2}. |
Denote
U1={(x,y)∈R2+|0<x≤ε1},U2={(x,y)∈R2+|0<y≤ε2}, |
U3={(x,y)∈R2+|x≥1ε1},U4={(x,y)∈R2+|ε1<x≤1ε1,y≥1ε2}. |
Evidently, Ucε1,2=U1∪U2∪U3∪U4. Now we just need to demonstrate that L˜V(x,y) is negative for any (x,y)∈Ucε1,2.
Case1. When (x,y)∈U1, in the light of (3.5), we can get
L˜V(x,y)≤−bx2+(a+b)x−kxyr1+x2−1y(c−β2)−λ≤(a+b)ε1−λ≤−λ2. |
Case2. When (x,y)∈U2, in consideration of (3.6), it is obvious that
L˜V(x,y)≤−bx2+(a+b)x−kxyr1+x2−1y(c−β2)−λ≤−1y(c−β2)−bx2+(a+b)x≤−1ε2(c−β2)+(a+b)24b≤−1. |
Case3. When (x,y)∈U3, according to (3.7), we can see that
L˜V(x,y)≤−bx2+(a+b)x−kxyr1+x2−1y(c−β2)−λ≤−b2x2−b2x2+(a+b)x≤−b2ε12+(a+b)22b≤−1. |
Case4. When (x,y)∈U4, in view of (3.8) to have
L˜V(x,y)≤−bx2+(a+b)x−kxyr1+x2−1y(c−β2)−λ≤(a+b)x−kxyr1+x2−λ≤x[(a+b)−kε12ε2(r1ε12+1)]−λ≤−λ. |
Hence, L˜V(x,y) is negative for any (x,y)∈Ucε1,2, the second condition is proved. Apparently, we have verified the conclusion of Theorem 3.3.
In this paper, a modified Leslie-Gower predation model is established, and three ecological issues are considered. First, the functional response of predator to prey is a rational non-monotone function. The second is that the predator is omnivore and will look for alternative food when the preferred food is scarce. Third, the intrinsic growth rates of the populations are disturbed by white noise. Coupling the environmental white noise into the differential equation for modeling is more consistent with the survival environment of the population. By comparing the deterministic system with the corresponding stochastic system, we obtain rich and interesting dynamic properties. To our knowledge, there are few studies that consider the effects of Holling-type IV functional response, generalist predators and environmental noise disturbance to model simultaneously.
First of all, we testify that the deterministic system (1.3) is permanent. Then we prove the stability of the positive equilibrium point E∗ by applying the Routh-Hurwitz criterion and constructing a suitable Lyapunov function. Next, we use numerical simulations to illustrate the stability of E∗. We choose parameters:
a=4,b=0.8,k=0.7,r1=4,c=0.2,p=0.4,r2=1. |
By calculating we can know that (H1) is satisfied, which means that the system (1.3) has the positive equilibrium point E∗. Through further calculation, the condition (H3) and (H4) are also satisfied. Then E∗=(4.91,2.95) is globally asymptotically stable. Figure 1 is obtained by numerical simulations, which describes the trend of solution with the initial value (0.2,0.1) (see Figure 1).
In addition, we research the dynamic behavior of the stochastic system (1.4), which takes into account the influence of white noise. To start with, we prove the existence and uniqueness of global positive solution. Secondly, we discover the conditions that make the population persistent in mean or extinct in four cases. And we conclude that when the white noise intensity is small enough, the population is persistent in mean. Nevertheless, when the intensity of white noise increases gradually, the population will be threatened and tend to be extinct. Furthermore, it can be seen that when the white noise of the stochastic system (1.4) is small, the population densities fluctuate around the deterministic steady-state values of x∗ and y∗. Next, we apply numerical simulations to validate our conclusions.
In order to satisfy the conditions of Theorem 3.2, we take α=3.4,β=1.5; α=0.05,β=1; α=3.3,β=0.04; α=0.05,β=0.05, respectively. Other parameters in model (1.4) have the same values as Figure 1. Using Matlab to get the following images, which conform to our theoretical results(see Figures 2–5).
On the other hand, there is a stationary distribution in system (1.4), which means that the two populations tend to coexist for a long time and the system is weakly stable. Similarly, the other parameters are the same as in Figure 1, we take α=0.06, β=0.04 to satisfy the conditions of Theorem 3.3, and we can see from the images that the densities of the two populations are close to the positive equilibrium point of x∗ and y∗ (see Figures 6–8).
As for the analysis of two systems, we can know that the difference between the stochastic and deterministic systems is that the stochastic system will inevitably be disturbed by the environment. To study the stochastic system is more in line with the characteristics of population environment variability. By comparing the two systems with numerical simulations, we can clearly see that when the white noise intensity is small, the population is persistent and fluctuates around the equilibrium state of the deterministic model. As the noise intensity increases, the population density oscillates and may eventually lead to extinction. Therefore, the existence of white noise is not conducive to the long-term survival of the population. From the point of view of pest management, white noise is a favorable factor and can be used for biological control. Obviously, the study of environmental noise disturbance is of great significance not only for integrated pest management but also for the development and utilization of biological resources.
This work is supported by Shandong Provincial Natural Science Foundation of China (No. ZR2019MA003).
The authors declare that there is no conflict of interests regarding the publication of this paper.
[1] |
S. C. Hoyt, Integrated chemical control of insects and biological control of mites on apple in washington, J. Econ. Entomol., 62 (1969), 74–86. doi: 10.1093/jee/62.1.74
![]() |
[2] |
D. J. Wollkind, J. B. Collings, J. A. Logan, Metastability in a temperature-dependent model system for predator-prey mite outbreak interactions on fruit trees, Bull. Math. Biol., 50 (1988), 379–409. doi: 10.1016/S0092-8240(88)90005-5
![]() |
[3] |
J. B. Collings, The effects of the functional response on the bifurcation behavior of a mite predator-prey interaction model, J. Math. Biol, 36 (1997), 149–168. doi: 10.1007/s002850050095
![]() |
[4] | R. J. Taylor, Predation, New York: Chapman and Hall, (1984). |
[5] | M. W.Sabelis, Predation on Spider Mites, Amsterdam: Elsevier, (1985), 103–129. |
[6] |
W. Sokol, J. A. Howell, Kinetics of phenol oxidation by washed cells, Biotechnol. Bioeng., 23 (1981), 2039–2049. doi: 10.1002/bit.260230909
![]() |
[7] | B. Gonzalez-Yanez, E. Gonzalez-Olivares, J. Mena-Lorca, Multistability on a Leslie-Gower type predator-prey model with nonmonotonic functional response, In R. Mondaini and R. Dilao (eds.), BIOMAT 2006 - International Symposium on Mathematical and Computational Biology, World Scienti.c Co. Pte. Ltd. (2007), 359–384. |
[8] |
Y. Li, D. Xiao, Bifurcations of a predator-prey system of Holling and Leslie types, Chaos Soliton. Fract., 34 (2007), 606–620. doi: 10.1016/j.chaos.2006.03.068
![]() |
[9] |
P. Ye, D. Wu, Impacts of strong Allee effect and hunting cooperation for a Leslie-Gower predator-prey system, Chin. J. Phys., 68 (2020), 49–64. doi: 10.1016/j.cjph.2020.07.021
![]() |
[10] | Y. Liu, Z. Guo, M. E. Smaily, L. Wang, A Leslie-Gower predator-prey model with a free boundary, Discret. Contin. Dyn. Syst., 12 (2019), 2063–2084. |
[11] |
M.A. Aziz-Alaoui, M. Daher Okiye, Boundedness and global stability for a predator-prey model with modified Leslie-Gower and Holling-type II schemes, Appl. Math. Lett., 16 (2003), 1069–1075. doi: 10.1016/S0893-9659(03)90096-6
![]() |
[12] | L. Puchuri Medina, E. Gonzalez-Olivares, A. Rojas-Palma, Multistability in a Leslie-Gower type predation model with rational nonmonotonic functional response and generalist predators, Comp. Math. Methods, 2 (2020). |
[13] | J. Liu, W. Zhao, The dynamic analysis of a stochastic prey-predator model with Markovian switching and different functional responses, Math. Model. Appl., 7 (2018), 12–21. |
[14] | T. Zhang, N. Gao, J. Wang, Z. Jiang, Dynamic system of microbial culture described by impulsive differential equations, Math. Model. Appl., 8 (2019), 1–13. |
[15] | X. Lv, X. Meng, Analysis of dynamic behavior for a nonlinear stochastic non-autonomous SIRS model, Math. Model. Appl., 7 (2018), 16–23. |
[16] |
P. Aguirre, E. Gonzalez-Olivares, S. Torres, Stochastic predator-prey model with Allee effect on preys, Nonlinear Anal.-Real World Appl., 14 (2013), 768–779. doi: 10.1016/j.nonrwa.2012.07.032
![]() |
[17] | X. Zhao, Z. Zeng, Stationary distribution of a stochastic predator-prey system with stage structure for prey, Phys. A, (2020), 545. |
[18] |
Q. Liu, D. Jiang, T. Hayat, A. Alsaedi, Dynamical behavior of stochastic predator-prey models with distributed delay and general functional response, Stoch. Anal. Appl., 38 (2020), 403–426. doi: 10.1080/07362994.2019.1695628
![]() |
[19] |
D. Manna, A. Maiti, G.P. Samanta, Deterministic and stochastic analysis of a predator-prey model with Allee effect and herd behaviour, Simulation., 95 (2019), 339–349. doi: 10.1177/0037549718779445
![]() |
[20] |
Q. Liu, D. Jiang, Influence of the fear factor on the dynamics of a stochastic predator-prey model, Appl. Math. Lett., 112 (2021), 106756. doi: 10.1016/j.aml.2020.106756
![]() |
[21] | C. Xu, Y. Yu, G. Ren, Dynamic analysis of a stochastic predator-prey model with Crowley-Martin functional response, disease in predator, and saturation incidence, J. Comput. Nonlinear Dyn., 15 (2020). |
[22] | V. Lakshmikantham, D. D. Bainov, P. S. Simeonov, Theory of impulsive differential equations, Singapore: World Scientific (1989), 265–273. |
[23] |
F. Chen, Z. Li, Y. Huang, Note on the permanence of a competitive system with infinite delay and feedback controls, Nonlinear Anal. RWA, 8 (2007), 680–687. doi: 10.1016/j.nonrwa.2006.02.006
![]() |
[24] | L. Arnold, Stochastic differential equations: Theory and applications, Bull. London Math. Soc., 8 (1976), 409. |
[25] |
P. S. Mandal, M. Banerjee, Stochastic persistence and stationary distribution in a Holling-Tanner type prey-predator model, Phys. A, 391 (2012), 1216–1233. doi: 10.1016/j.physa.2011.10.019
![]() |
[26] |
C. Ji, D. Jiang, N. Shi, Analysis of a predator-prey model with modified Leslie-Gower and Holling-type II schemes with stochastic perturbation, J. Math. Anal. Appl., 359 (2009), 482–498. doi: 10.1016/j.jmaa.2009.05.039
![]() |
[27] | D. Zhou, M. Liu, Z. Liu, Persistence and extinction of a stochastic predator-prey model with modified Leslie-Gower and Holling-type II schemes, Adv. Differ. Equ., 2020 (2020). |
[28] | D. Xu, M. Liu, X. Xu, Analysis of a stochastic predator-prey system with modified Leslie-Gower and Holling-type IV schemes, Phys. A, 537 (2019), 122761. |
[29] |
H. Li, F. Cong, Dynamics of a stochastic Holling-Tanner predator-prey model, Phys. A, 531 (2019), 121761. doi: 10.1016/j.physa.2019.121761
![]() |
1. | Yousef Alnafisah, Moustafa El-Shahed, Deterministic and stochastic model for the hepatitis C with different types of virus genome, 2022, 7, 2473-6988, 11905, 10.3934/math.2022664 | |
2. | Yousef Alnafisah, Moustafa El-Shahed, Stochastic Analysis of a Hantavirus Infection Model, 2022, 10, 2227-7390, 3756, 10.3390/math10203756 | |
3. | Yudan Ma, Ming Zhao, Yunfei Du, Impact of the strong Allee effect in a predator-prey model, 2022, 7, 2473-6988, 16296, 10.3934/math.2022890 | |
4. | Moustafa El-Shahed, Asma Al-Nujiban, Nagdy F. Abdel-Baky, Stochastic Modelling of Red Palm Weevil Using Chemical Injection and Pheromone Traps, 2022, 11, 2075-1680, 334, 10.3390/axioms11070334 | |
5. | Shivani Khare, Kunwer Singh Mathur, Rajkumar Gangele, 2022, Chapter 96, 978-3-030-99791-5, 1133, 10.1007/978-3-030-99792-2_96 | |
6. | Daliparthi Bhanu Prakash, Dasu Krishna Kiran Vamsi, Stochastic optimal and time-optimal control studies for additional food provided prey–predator systems involving Holling type III functional response, 2023, 11, 2544-7297, 10.1515/cmb-2022-0144 | |
7. | D. Bhanu Prakash, D. K. K. Vamsi, Stochastic time-optimal control and sensitivity studies for additional food provided prey-predator systems involving Holling type-IV functional response, 2023, 9, 2297-4687, 10.3389/fams.2023.1122107 |