
In this paper, a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps is studied. Firstly, it is proved that the model has a unique global solution. Secondly, sufficient conditions for weak persistence in the mean and extinction of prey and predator populations are established. Finally, sufficient conditions for the existence and uniqueness of ergodic stationary distribution are established. Moreover, several numerical simulations are presented to illustrate the main results.
Citation: Hong Qiu, Yanzhang Huo, Tianhui Ma. Dynamical analysis of a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps[J]. AIMS Mathematics, 2022, 7(8): 14492-14512. doi: 10.3934/math.2022799
[1] | Yuanfu Shao . Dynamics and optimal harvesting of a stochastic predator-prey system with regime switching, S-type distributed time delays and Lévy jumps. AIMS Mathematics, 2022, 7(3): 4068-4093. doi: 10.3934/math.2022225 |
[2] | Xuegui Zhang, Yuanfu Shao . Analysis of a stochastic predator-prey system with mixed functional responses and Lévy jumps. AIMS Mathematics, 2021, 6(5): 4404-4427. doi: 10.3934/math.2021261 |
[3] | 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 |
[4] | 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 |
[5] | Francesca Acotto, Ezio Venturino . How do predator interference, prey herding and their possible retaliation affect prey-predator coexistence?. AIMS Mathematics, 2024, 9(7): 17122-17145. doi: 10.3934/math.2024831 |
[6] | Weili Kong, Yuanfu Shao . The effects of fear and delay on a predator-prey model with Crowley-Martin functional response and stage structure for predator. AIMS Mathematics, 2023, 8(12): 29260-29289. doi: 10.3934/math.20231498 |
[7] | Xiaodong Wang, Kai Wang, Zhidong Teng . Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence. AIMS Mathematics, 2023, 8(2): 2829-2855. doi: 10.3934/math.2023148 |
[8] | Jorge Luis Ramos-Castellano, Miguel Angel Dela-Rosa, Iván Loreto-Hernández . Bifurcation analysis for the coexistence in a Gause-type four-species food web model with general functional responses. AIMS Mathematics, 2024, 9(11): 30263-30297. doi: 10.3934/math.20241461 |
[9] | Yassine Sabbar, Aeshah A. Raezah . Modeling mosquito-borne disease dynamics via stochastic differential equations and generalized tempered stable distribution. AIMS Mathematics, 2024, 9(8): 22454-22485. doi: 10.3934/math.20241092 |
[10] | Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045 |
In this paper, a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps is studied. Firstly, it is proved that the model has a unique global solution. Secondly, sufficient conditions for weak persistence in the mean and extinction of prey and predator populations are established. Finally, sufficient conditions for the existence and uniqueness of ergodic stationary distribution are established. Moreover, several numerical simulations are presented to illustrate the main results.
In ecological mathematics, in order to simulate a variety of different situations (see, e.g., [1,2]), many different kinds of predator-prey models [3,4,5,6] are studied. In 1975, Beddington-DeAngelis [7,8] showed that the predator-dependent functional response can provide a better description for predator-prey models, and introduced the following Beddington-DeAngelis type predator-prey model:
{dx(t)=x(t)(r1−a11x(t)−a12y(t)1+mx(t)+ny(t))dt,dy(t)=y(t)(−r2+a21x(t)1+mx(t)+ny(t))dt, | (1.1) |
where x(t) denotes the population densities of prey at time t, and y(t) represents the population densities of predator at time t. r1 denotes the intrinsic growth rate of the prey, r2 denotes the mortality rate of the predator, a11 represents the density dependent coefficient of prey, a12 represents the capturing rate of predator, a21 represents the rate at which nutrients are converted into reproduction of the predator. In addition, ri,aij,m,n are all positive constants, i,j=1,2.
Species are inevitably disturbed by environmental noise (see, e.g., [9,10,11,12,13,14,15]). Therefore, it is of great significance to study the effect of noise on biological population system. A large number of authors have studied stochastic species models with random factors. Du et al. investigated dynamics of a stochastic Lotka-Volterra model perturbed by white noise [10]. Mao et al. got that even small enough noise can have an effect for explosion in species dynamics in [12]. From [14], we know that Wang et al. studied stationary distribution of a stochastic hybrid phytoplankton-zooplankton model with toxin-producing phytoplankton. Considering the influence of random environment, random disturbances should be introduced into the model to explore the effects of random disturbances. On the one hand, it is assumed that the random interference is white noise. Therefore, we obtain the following stochastic Beddington-DeAngelis type predator-prey model with white noise:
{dx(t)=x(t)(r1−a11x(t)−a12y(t)1+mx(t)+ny(t))dt+σ1x(t)dB1(t),dy(t)=y(t)(−r2−a22y(t)+a21x(t)1+mx(t)+ny(t))dt+σ2y(t)dB2(t), | (1.2) |
where σ2i(i=1,2) stands for the intensity of the white noise, {B1(t),B2(t)}t≥0 is a two dimensional Brownian motion. Throughout this paper, the Brownian motion was defined on a complete probability space (Ω,{Ft}t≥0,P) with a filtration {Ft}t∈R+ satisfying the usual conditions.
Because growth rates and death rates are especial sensitive to the disturbance of the environment. It is known that the rates often switch randomly because of changes in the environment. For instance, temperature or rain falls [16]. A classic example [17] is that some species grow at vastly different rates during the dry and rainy seasons. It has been suggested [16,17,18,19] that these random changes can be simulated using a continuous time finite state Markov chains. Hence, considering this influence, we get the following model:
{dx(t)=x(t)[r1(λ(t))−a11x(t)−a12y(t)1+mx(t)+ny(t)]dt+σ1(λ(t))x(t)dB1(t),dy(t)=y(t)[−r2(λ(t))−a22y(t)+a21x(t)1+mx(t)+ny(t)]dt+σ2(λ(t))y(t)dB2(t), | (1.3) |
where λ(t) represents a continuous-time Markov chain with a state space S={1,2,3,⋯,n}.
On the other hand, white noise is not a good description for some sudden environmental disturbances (such as earthquakes, volcanoes, hurricanes, debris flows, etc.) that are often encountered during the growth of species. Bao et al. [20,21] suggested that these phenomena can be described by Lévy jumps, and they studied the stochastic Lotka-Volterra population systems with Lévy jumps. In 2015, Zang et al. [22] studied dynamics of a stochastic predator-prey model with Beddington-DeAngelis functional response and Lévy jumps. To the best of our knowledge, no results related to Beddington-DeAngelis predator-prey model with regime-switching and Lévy jumps has been reported. Motivated by this, we will consider the following model:
{dx(t)=x(t−)[r1(λ(t))−a11x(t)−a12y(t)1+mx(t)+ny(t)]dt+σ1(λ(t))x(t−)dB1(t)+∫Zx(t−)γ1(λ(t),u)˜N(dt,du),dy(t)=y(t−)[−r2(λ(t))−a22y(t)+a21x(t)1+mx(t)+ny(t)]dt+σ2(λ(t))y(t−)dB2(t)+∫Zy(t−)γ2(λ(t),u)˜N(dt,du), | (1.4) |
where x(t−) and y(t−) are the left limit of x(t) and y(t), respectively. ˜N(dt,du)=N(dt,du)−λ(du)dt, on a measurable subset Z of (0,+∞) with λ(Z)<+∞, N is a Poisson counting measure with characteristic measure λ, γi:Z×Ω→R is bounded and continuous with respect to λ, i=1,2. Let's make the following assumption
1+γi(u)>0,u∈Z,i=1,2. | (1.5) |
To better simulate the population, a predat-prey model with Bedtington-DeAngelis functional response is studied in this paper. In our model, predator not only has time to find the prey, but also has time to deal with the encounter with other predators. We also consider the influence of environmental factors. The value of this paper is to study the stochastic predator-prey model of Bedtington-DeAngelis function response which is affected by both regime-switching and Lévy jumps. Furthermore, we introduce some numerical simulations to illustrate the main results.
For the sake of convenience, we define some notations.
R2+={m∈R2∣mi>0,i=1,2}, |
Φ1(t)=∫t0σ1(λ(s))dB1(s),Φ2(t)=∫t0∫Zln(1+γ1(λ(s),u))˜N(ds,du), |
Φ3(t)=∫t0σ2(λ(s))dB2(s),Φ4(t)=∫t0∫Zln(1+γ2(λ(s),u))˜N(ds,du), |
b1(j)=r1(j)−12σ21(j)−∫Z[γ1(j,u)−ln(1+γ1(j,u))]λ(du),ˉb1=∑j∈Sπjb1(j), |
b2(j)=−r2(j)−12σ22(j)−∫Z[γ2(j,u)−ln(1+γ2(j,u))]λ(du),ˉb2=∑j∈Sπjb2(j). |
Assumption 2.1. For any j∈S, assume that there exists a constant G1>0 such that
∫Z{|γi(j,u)|2∨[ln(1+γi(j,u))]2}λ(du)≤G1<+∞,i=1,2. |
That is to say, the intensity of Lévy noise is not too large.
Assumption 2.2. For any j∈S, assume that there exist two constants G2,G3>0 such that
∫Z{(1+γi(j,u))p−1−pγi(j,u)}λ(du)≤G2<+∞, |
∫Z[(γi(j,u)−ln(1+γi(j,u))]λ(du)≤G3<+∞,i=1,2. |
Lemma 2.1. Assumptions 2.1 and 2.2 hold, then the model (1.4) with initial value (x(0),y(0),λ(0)) has a unique positive local solution (x(t),y(t),λ(t)) on t∈[0,τe) almost surely(a.s.), where τe is the explosion time.
Proof. Consider the following equations:
{dφ(t)=[c1(λ(t))−a11eφ(t)−a12eψ(t)1+meφ(t)+neψ(t)]dt+σ1(λ(t))dB1(t)+∫Zγ1(λ(t),u)˜N(dt,du),dψ(t)=[−c2(λ(t))−a22eψ(t)+a21eφ(t)1+meφ(t)+neψ(t)]dt+σ2(λ(t))dB2(t)+∫Zγ2(λ(t),u)˜N(dt,du), | (2.1) |
with initial value (φ(0)=lnx(0),ψ(0)=lny(0),λ(0)), where
c1(λ(t))=r1(λ(t))−12σ21(λ(t))+∫Z[ln(1+γ1(λ(t),u))−γ1(λ(t),u)]du, |
c2(λ(t))=r2(λ(t))+12σ22(λ(t))−∫Z[ln(1+γ2(λ(t),u))−γ2(λ(t),u)]du. |
According to the Lemma 1 of reference [23], we know that the coefficients of model (2.1) are local Lipschitz continuous, hence there is a unique local solution (φ(t),ψ(t),λ(t)) on t∈[0,τe). That is to say, (x(t)=eφ(t),y(t)=eψ(t),λ(t)) is the unique positive local solution of (1.4) with initial value (x(0),y(0),λ(0)) by Itˆo's formula.
Theorem 2.1. Consider the model (1.4), for given initial value(X(0),λ(0))∈R2+×S, where X(0)=(x(0),y(0)), there is a unique global solution (x(t),y(t),λ(t)) on t≥0, and the solution will remain in R2+ with probability 1.
Proof. By Lemma 2.1, we just need to prove τe=+∞. Consider the following equations:
dM1(t)=M1(t−)[r1(λ(t))−a11M1(t)]dt+σ1(λ(t))M1(t−)dB1(t)+∫ZM1(t−)γ1(λ(t),u)˜N(dt,du),M1(0)=x(0). | (2.2) |
dM2(t)=M2(t−)[−r2(λ(t))−a22M2(t)]dt+σ2(λ(t))M2(t−)dB2(t)+∫ZM2(t−)γ2(λ(t),u)˜N(dt,du),M2(0)=y(0). | (2.3) |
dM3(t)=M3(t−)[−r2(λ(t))−a22M3(t)+a21M1(t)]dt+σ2(λ(t))M3(t−)dB2(t)+∫ZM3(t−)γ2(λ(t),u)˜N(dt,du),M3(0)=y(0). | (2.4) |
By the comparison theorem of stochastic equation (see e.g., [24] (Theorem 1 on page 173) or [25] (Theorem 3.1)), we get
x(t)≤M1(t),M2(t)≤y(t)≤M3(t),t∈[0,τe). |
According to Lemma 3.1 of reference [26], (2.2)–(2.4) have the following explicit solutions:
M1(t)=exp{∫t0(r1(λ(s))−β1(λ(s)))ds+∫t0σ1(λ(s))dB1(s)+k1(t)}x−1(0)+a11∫t0exp{∫s0(r1(λ(τ))−β1(λ(τ)))dτ+∫s0σ1(λ(τ))dB1(τ)+k1(s)}ds, |
M2(t)=exp{∫t0(−r2(λ(s))−β2(λ(s)))ds+∫t0σ2(λ(s))dB2(s)+k2(t)}y−1(0)+a22∫t0exp{∫s0(−r2(λ(τ))−β2(λ(τ)))dτ+∫s0σ2(λ(τ))dB2(τ)+k2(s)}ds, |
![]() |
where
βi(λ(t))=12σ2i(λ(t))+∫Z[γi(λ(t),u)−ln(1+γi(λ(t),u))]λ(du),i=1,2, |
ki(t)=∫t0∫Zln(1+γi(λ(s),u))˜N(ds,du),i=1,2. |
Because M1(t),M2(t) and M3(t) are existent on t≥0, then we obtain τe=+∞.
Lemma 2.2. (Lemma 8 in [27]) Suppose that x(t)∈C(Ω×[0,+∞),R+), and limt→+∞F(t)t=0 a.s.
(1) If there exist two constants T>0 and ρ0>0 such that for all t≥T,
lnx(t)≤ρt−ρ0∫t0x(s)ds+F(t), |
then
{lim supt→+∞t−1∫t0x(s)ds≤ρρ0a.s.,ifρ≥0;limt→+∞x(t)=0a.s.,ifρ<0. |
(2) If there exist three constants T>0, ρ>0 and ρ0>0 such that for all t≥T,
lnx(t)≥ρt−ρ0∫t0x(s)ds+F(t), |
then lim inft→+∞t−1∫t0x(s)ds≥ρρ0 a.s.
Lemma 2.3. The solution of system (1.4) have the following properties:
lim supt→+∞lnx(t)t≤0,lim supt→+∞lny(t)t≤0,a.s.. |
Proof. From system (1.4), we get
{dx(t)≤x(t−)[r1(λ(t))−a11x(t)]dt+σ1(λ(t))x(t−)dB1(t)+∫Zx(t−)γ1(λ(t),u)˜N(dt,du),dy(t)≤y(t−)[a21m−r2(λ(t))−a22y(t)]dt+σ2(λ(t))y(t−)dB2(t)+∫Zy(t−)γ2(λ(t),u)˜N(dt,du). |
Set
{dˉx(t)=ˉx(t−)[r1(λ(t))−a11ˉx(t)]dt+σ1(λ(t))ˉx(t−)dB1(t)+∫Zˉx(t−)γ1(λ(t),u)˜N(dt,du),dˉy(t)=ˉy(t−)[a21m−r2(λ(t))−a22ˉy(t)]dt+σ2(λ(t))ˉy(t−)dB2(t)+∫Zˉy(t−)γ2(λ(t),u)˜N(dt,du), | (2.5) |
where (ˉx(t),ˉy(t),λ(t)) is a solution of model (2.5) with initial value (x(0),y(0),λ(0)). According to the comparison theorem of stochastic differential equations, it's easy to figure out x(t)≤ˉx(t) for t≥0.
By using Lemma 2 of reference [28], one obtain the following results for t≥0: lim supt→+∞lnˉx(t)lnt≤1, lim supt→+∞lnˉy(t)lnt≤1,a.s., therefore, we get lim supt→+∞lnx(t)lnt≤1. That is to say, we have lim supt→+∞lnx(t)t=lim supt→+∞lnx(t)lntlim supt→+∞lntt≤lim supt→+∞lntt=0. Similarly, we have lim supt→+∞lny(t)t≤0.
Definition 2.1. [29] The population x(t) is weakly persistent in the mean if ⟨x⟩∗>0, where ⟨x(t)⟩=1t∫t0x(s)ds, (x(t))∗=lim supt→+∞x(t).
Theorem 2.2. Assumptions 2.1 and 2.2 hold.
(i) The prey x(t) will go to extinction if ˉb1<0, i.e., limt→+∞x(t)=0a.s..
(ii) The prey x(t) will be weakly persistent in the mean if ˉb1>0, i.e., ⟨x(t)⟩∗>0a.s..
Proof. (i) Applying Itˆo's formula to system (1.4), we obtain
{dlnx(t)=[b1(λ(t))−a11x(t)−a12y(t)1+mx(t)+ny(t)]dt+σ1(λ(t))dB1(t)+∫Z(1+γ1(λ(t),u))˜N(dt,du),dlny(t)=[b2(λ(t))−a22y(t)+a21x(t)1+mx(t)+ny(t)]dt+σ2(λ(t))dB2(t)+∫Z(1+γ2(λ(t),u))˜N(dt,du), | (2.6) |
where
b1(λ(t))=r1(λ(t))−12σ21(λ(t))−∫Z[γ1(λ(t),u)−ln(1+γ1(λ(t),u))]du, |
b2(λ(t))=−r2(λ(t))−12σ22(λ(t))−∫Z[γ2(λ(t),u)−ln(1+γ2(λ(t),u))]du. |
By the ergodicity of the Markov chain, we have
limt→+∞t−1∫t0bi(λ(s))ds=ˉbi,i=1,2. | (2.7) |
Hence we have ˉb2<0. From the first equation of (2.6), we get
lnx(t)−lnx(0)=∫t0[b1(λ(s))−a11x(s)−a12y(s)1+mx(s)+ny(s)]ds+∫t0σ1(λ(s))dB1(s)+∫t0∫Zln(1+γ1(λ(s),u))˜N(ds,du). |
so we have
lnx(t)−lnx(0)t≤⟨b1⟩+Φ1(t)t+Φ2(t)t. | (2.8) |
According to Assumption 2.1 and the strong law of large numbers [30], we can see that
limt→+∞t−1Φi(t)=0,i=1,2,3,4, | (2.9) |
taking superior limit for both sides of (2.8), one obtain lim supt→+∞lnx(t)t≤ˉb1<0. Namely, limt→+∞x(t)=0.
(ii) We just need to prove that there is a constant u1>0 that makes ⟨x(t)⟩∗=u1>0a.s. true for the solution (x(t),y(t),λ(t)) of system (1.4) that has the initial value (x(0),y(0),λ(0)). If not, for any ε1>0, there exists a solution (x1(t),y1(t),λ(t)) with an initial value (x(0),y(0),λ(0)) such that P{⟨x1(t)⟩∗<ε1}>0 holds. Suppose ε1 is small enough, and satisfy the following conditions:
ˉb1−a11ε1>0,ˉb2+a21ε1<0. |
According to the second equation of (2.6), we have
lny1(t)−lny(0)t≤⟨b2⟩−a22⟨y1(t)⟩+a21⟨x1(t)⟩+Φ3(t)t+Φ4(t)t, | (2.10) |
taking superior limit on both sides of (2.10), using virtue of (2.7), (2.9) and (2.10), we have
lim supt→+∞t−1lny1(t)≤ˉb2+a21ε1<0. |
So, we obtain
limt→+∞y1(t)=0. | (2.11) |
By using the first equation of (2.6), we have
lnx1(t)−lnx(0)=∫t0[b1(λ(s))−a11x1(t)−a12y1(t)1+mx1(t)+ny1(t)]ds+∫t0σ1(λ(s))dB1(s)+∫t0∫Zln(1+γ1(λ(s),u))˜N(ds,du), |
therefore, we obtain
lnx1(t)−lnx(0)t≥⟨b1⟩−a11⟨x1(t)⟩−a12⟨y1(t)⟩+Φ1(t)t+Φ2(t)t. | (2.12) |
By the same method, using virtue of (2.7), (2.9), (2.11) and (2.12), one obtain
(t−1lnx1(t))∗≥ˉb1−a11ε1>0. |
Namely, one have P{(t−1lnx1(t))∗>0}>0, this is a contradiction to Lemma 2.3. So ⟨x(t)⟩∗>0, x(t) will be weakly persistent in the mean a.s.. Complete the proof.
Theorem 2.3. Assumptions 2.1 and 2.2 hold.
(i) The predator y(t) will go to extinction if a11ˉb2+a21ˉb1<0, i.e., limt→+∞y(t)=0a.s..
(ii) The predator y(t) will be weakly persistent in the mean if ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗>0, i.e., ⟨y(t)⟩∗>0a.s., where (ˉx,ˉy) is a solution of system (2.5) with initial value (x(0),y(0),λ(0)).
Proof. (i) If ˉb1≤0, from the Theorem 2.2, we can know that limt→∞x(t)=0. According to (2.10), we have
lny(t)−lny(0)t≤⟨b2⟩+a21⟨x(t)⟩+Φ3(t)t+Φ4(t)t. |
So we obtain lim supt→+∞t−1lny(t)≤ˉb2<0, then limt→+∞y(t)=0.
If ˉb1>0, according to the property of limit and (2.9), there exists a T for sufficiently small ε such that
lnx(t)−lnx(0)t≤ˉb1−a11⟨x(t)⟩+Φ1(t)t+Φ2(t)t≤ˉb1−a11⟨x(t)⟩+ε, |
on t>T. Using Lemma 2.2 and the arbitrariness of ε, one can obtain ⟨x(t)⟩∗≤ˉb1a11. Substituting the inequality into the second equation of (2.6). By further calculation, we have
lim supt→+∞t−1lny(t)≤ˉb2+⟨a21x(t)⟩∗≤ˉb2+a21ˉb1a11<0. |
Namely, we get limt→+∞y(t)=0 a.s..
(ii) We just need to prove that there is a constant u2>0 that makes ⟨y(t)⟩∗=u2>0a.s. true for the solution (x(t),y(t),λ(t)) of system (1.4) that has the initial value (x(0),y(0),λ(0)). If not, for any ε2>0, there exists a solution (x2(t),y2(t),λ(t)) with an initial value (x(0),y(0),λ(0)) such that P{⟨y2(t)⟩∗<ε2}>0 holds. Suppose ε2 is small enough, and satisfy the following conditions:
ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗−(a22+2a12a21a11)ε2>0. |
From the second equation of (2.6), one can get
lny2(t)−lny(0)t=⟨b2⟩+⟨a21ˉx1+mˉx+nˉy⟩−⟨a22y2(t)⟩+Φ3(t)t+Φ4(t)t+⟨a21x21+mx2+ny2−a21ˉx1+mˉx+nˉy⟩, | (2.13) |
where (ˉx,ˉy) is a solution of system (2.5), we have x2(t)≤ˉx(t), y2(t)≤ˉy(t),a.s. on t>0.
Due to
a21x21+mx2+ny2−a21ˉx1+mˉx+nˉy=a21nˉx(ˉy−y2)−a21(ˉx−x2)−a21nˉy(ˉx−x2)[1+mx2+ny2][1+mˉx+nˉy]≥a21nˉx(ˉy−y2)[1+mx2+ny2][1+mˉx+nˉy]−a21(ˉx−x2)−a21nˉy(t)(ˉx−x2)nˉy(t)≥−2a21(ˉx−x2(t)), |
thus we have
lny2(t)−lny(0)t≥⟨b2⟩+⟨a21ˉx1+mˉx+nˉy⟩−⟨a22y2(t)⟩+Φ3(t)t+Φ4(t)t−⟨2a21(ˉx−x2(t))⟩. | (2.14) |
Next, we define the function V1(t):=|lnˉx(t)−lnx2(t)|, by using Itˆo's formula, we obtain
dV1(t)=[−a11(ˉx(t)−x2(t))+a12y2(t)1+mx2(t)+ny2(t)]dt≤[a12y2(t)−a11(ˉx(t)−x2(t))]dt, |
integrating and dividing by t on both sides of the above inequality, one obtain V1(t)−V1(0)t≤a12⟨y2(t)⟩−a11⟨ˉx(t)−x2(t)⟩. Since V1(t)t≥0, we have a11⟨ˉx(t)−x2(t)⟩≤a12⟨y2(t)⟩+V1(0)t, Note that V1(0)=0, thus ⟨ˉx(t)−x2(t)⟩≤a12a11⟨y2(t)⟩.
Substituting into (2.14), we get
lny2(t)−lny(0)t≥⟨b2⟩+⟨a21ˉx1+mˉx+nˉy⟩−⟨a22y2(t)⟩+Φ3(t)t+Φ4(t)t−2a21a12a11⟨y2(t)⟩. |
That is to say, we have
(t−1lny2(t))∗≥ˉb2+⟨a21ˉx1+mˉx+nˉy⟩∗−(a22+2a21a12a11)ε2>0, |
which contradicts Lemma 2.3, so we get ⟨y(t)⟩∗>0 a.s., i.e., y(t) is weakly persistent in the mean.
Lemma 2.4. ([31] Theorem 2.3) Let Assumptions 2.1 and 2.2 hold. For any q>0,t≥0, there exist constants Ki(q)>0,i=1,2 such that
supt∈R+E|x(t)|q≤K1(q),supt∈R+E|y(t)|q≤K2(q). |
Assumption 2.3. a11≥a12mn+2a21, a22≥a21nm+2a12.
Lemma 2.5. Let X((x0,y0,j),t)=(x((x0,y0,j),t),y((x0,y0,j),t)) is a solution of model (1.4) with initial data ((x0,y0),j)∈D×S, X((˜x0,˜y0,˜j),t)=(x((˜x0,˜y0,˜j),t),y((˜x0,˜y0,˜j),t)) is a solution of model (1.4) with initial data ((˜x0,˜y0),˜j)∈D×S respectively, where D is an arbitrary compact subset of R2+. If Assumptions 2.1–2.3 hold, then
limt→+∞(E|x((x0,y0,j),t)−x((˜x0,˜y0,˜j),t)|+E|y((x0,y0,j),t)−y((˜x0,˜y0,˜j),t)|)=0,a.s. |
Proof. For the sake of convenience, we define the following notations:
x=x((x0,y0,j),t),˜x=x((˜x0,˜y0,˜j),t),y=y((x0,y0,j),t),˜y=y((˜x0,˜y0,˜j),t). |
Define V2(t)=|lnx−ln˜x|+|lny−ln˜y|, then we have
d+V2(t)=sgn(x−˜x)d(lnx−ln˜x)+sgn(y−˜y)d(lny−ln˜y)=sgn(x−˜x){−a11(x−˜x)−a12(y−˜y)+m˜x(y−˜y)−m˜y(x−˜x)[1+mx+ny][1+m˜x+n˜y]}dt+sgn(y−˜y){−a22(y−˜y)+a21(x−˜x)+n˜y(x−˜x)−n˜x(y−˜y)[1+mx+ny][1+m˜x+n˜y]}dt≤{−a11|x−˜x|+2a12|y−˜y|+a12mn|x−˜x|}dt+{−a22|y−˜y|+2a21|x−˜x|+a21nm|y−˜y|}dt≤{−(a11−a12mn−2a21)|x−˜x|−(a22−2a12−a21nm)|y−˜y|}dt, |
therefore, we get
EV2(t)≤V2(0)−(a11−a12mn−2a21)∫t0E|x−˜x|ds−(a22−2a12−a21nm)∫t0E|y−˜y|ds. |
Note that EV2(t)≥0, so we have
(a11−a12mn−2a21)∫t0E|x−˜x|ds+(a22−2a12−a21nm)∫t0E|y−˜y|ds≤V2(0)<+∞. |
Hence E|x−˜x|∈L1[0,+∞),E|y−˜y|∈L1[0,+∞). By Barbˇa lat's conclusions (see, e.g., [32]), now we just proof that E(x(t)) and E(y(t)) are uniformly continuous with respect to t. As a matter of fact, thanks to system (1.4), one get
E(x(t))=x(0)+∫t0[E(r1(λ(s))x(s))−E(a11x2(s))−E(a12x(s)y(s)1+mx(s)+ny(s))]ds, |
E(y(t))=y(0)+∫t0[E(−r2(λ(s))y(s))−E(a22y2(s))+E(a21x(s)y(s)1+mx(s)+ny(s))]ds. |
So, E(x(t)) and E(y(t)) are continuously differentiable. In addition, according to above formulas and Lemma 2.4, we have
dE(x(t))dt≤E(x(t))ru1≤K1ru1,dE(y(t))dt≤E(y(t))a21m≤K2a21m, |
where K1>0,K2>0 are two constants, ru1=maxi∈Sr1(i). Namely, E(x(t)) and E(y(t)) are uniformly continuous.
Theorem 2.4. Suppose that Assumptions 2.1–2.3 hold, if ˉb1>0, ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗>0, then system (1.4) has a unique stationary measure η(⋅×⋅) which is ergodic.
To prove this theorem we need to introduce more notations. Let B(R2+×S) represent all the probability measures defined on R2+×S. For any two measures p1,p2∈B, define the metric dH as follows
dH(p1,p2)=suph∈H|n∑i=1∫R2+h(x,i)p1(dx,i)−n∑i=1∫R2+h(x,i)p2(dx,i)|, |
where
H={h:R2+×S→R||h(x,i)−h(y,j)|≤|x−y|+|i−j|,|h(⋅,⋅)|≤1}. |
Let us now present the following lemmas.
Lemma 2.6. For every q>0 and any compact subset D of R2+, sup(X(0),j)∈D×SE[sup0≤s≤t|XX(0),j(s)|q]<+∞,∀t≥0, where XX(0),j=(xX(0),j,yX(0),j).
Proof. From (1.4), we have
x(t)=x(0)+∫t0[x(s)r1(λ(s))−a11x2(s)−a12x(s)y(s)1+mx(s)+ny(s)]ds+∫t0σ1(λ(s))x(s)dB1(s)+∫t0∫Zx(s)γ1(λ(s),u)˜N(ds,du), |
y(t)=y(0)+∫t0[−y(s)r2(λ(s))−a22y2(s)+a21x(s)y(s)1+mx(s)+ny(s)]ds+∫t0σ2(λ(s))y(s)dB2(s)+∫t0∫Zy(s)γ2(λ(s),u)˜N(ds,du). |
According to the H¨older inequality and the moment inequality of stochastic integrals, exists k=1,2,⋅⋅⋅, such that
E[sup(k−1)ξ≤s≤kξ|x(s)|q]≤4q−1|x((k−1)ξ)|q+4q−1E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξ[x(s)r1(λ(s))−a11x2(s)−a12x(s)y(s)1+mx(s)+ny(s)]ds|q)+4q−1E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξσ1(λ(s))x(s)dB1(s)|q)+4q−1E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξ∫Zx(s)γ1(λ(s),u)˜N(ds,du)|q). | (2.15) |
By Lemma 2.4, exist a positive constant K(q) such that E|x(t)|q≤K(q), t∈[0,+∞). Thus we can obtain that
E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξ[x(s)r1(λ(s))−a11x2(s)−a12x(s)y(s)1+mx(s)+ny(s)]ds|q)≤E[ξqsup(k−1)ξ≤s≤kξ(|x(s)|q|r1(λ(s))−a11x(s)−a12y(s)1+mx(s)+ny(s)|q)]≤ξq⋅12E(|x(s)|2q)+12ξqE(sup(k−1)ξ≤s≤kξ|r1(λ(s))−a11x(s)−a12y(s)1+mx(s)+ny(s)|2q)≤12ξqE(|x(s)|2q)+12ξq⋅32q−1[(ru1)2q+a2q11E(sup(k−1)ξ≤s≤kξ|x(s)|2q)+(a12n)2q]:=M1(q)ξq. | (2.16) |
In view of the Burkholder-Davis-Gundy inequality (see Theorem 1.7.3 of reference [33]), we have
E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξσ1(λ(s))x(s)dB1(s)|q)≤CqE[∫kξ(k−1)ξ|x(s)σ1(λ(s))|2ds]q2≤Cqξq2(σu1)qE(sup(k−1)ξ≤s≤kξ|x(s)|q):=M2(q)ξq2, | (2.17) |
where σu1=maxi∈Sσ1(i). Make use of Assumption 2.1 and Kunita's first inequality (see Theorem 4.4.23 in reference [34]), one obtain
E(sup(k−1)ξ≤s≤kξ|∫kξ(k−1)ξ∫Zx(s)γ1(λ(s),u)˜N(ds,du)|q)≤Dq{E[∫kξ(k−1)ξ∫Z|x(s)γ1(λ(s),u)|2λ(du)ds]q2+E∫kξ(k−1)ξ∫Z|x(s)γ1(λ(s),u)|qλ(du)ds]}≤Dqξq2Gq21K(q)+DqξGq21K(q). | (2.18) |
According to (2.15)–(2.18), we have
sup(X(0),j)∈D×SE[sup0≤s≤t|xX(0),j(s)|q]<+∞,∀s∈[0,t],∀t≥0. |
Similarly, we obtain
sup(X(0),j)∈D×SE[sup0≤s≤t|yX(0),j(s)|q]<+∞,∀s∈[0,t],∀t≥0. |
Therefore, we have
sup(X(0),j)∈D×SE[sup0≤s≤t|XX(0),j(s)|q]<+∞,∀s∈[0,t],∀t≥0. |
Lemma 2.7. Assumption 2.1–2.3 hold, then for any compact subset D of R2+, we have
limt→+∞dH(p(t,X(0),i,⋅×⋅),p(t,˜X(0),j,⋅×⋅))=0 | (2.19) |
uniformly in X(0),˜X(0)∈D and i,j∈S.
Proof. The proof is similar to Lemma 5.6 of reference [13]. By using Lemmas 2.5 and 2.6, the proof is easy to prove, hence it is omitted.
Lemma 2.8. Assumption 2.1–2.3 hold. Then for any (X(0),i)∈R2+×S, {p(t,X(0),i,⋅×⋅)|t≥0} is Cauchy in the space B(R2+×S) with metric dH.
Proof. Fix any (X(0),i)∈R2+×S, we just need to proof that for any ε3>0, there is a T>0 such that
dH(p(t+s,X(0),i,⋅×⋅),p(t,X(0),i,⋅×⋅))≤ε3,∀t≥T,s>0. |
This is equivalent to
suph∈H|Eh(XX(0),i(t+s),λi(t+s))−Eh(XX(0),i(t),λi(t))|≤ε3,∀t≥T,s>0. | (2.20) |
For any h∈H and t,s>0, we have
|Eh(XX(0),i(t+s),λi(t+s))−Eh(XX(0),i(t),λi(t))|=|E[E(h(XX(0),i(t+s),λi(t+s))|Fs)]−Eh(XX(0),i(t),λi(t))|=|n∑l=1∫R2+Eh(Xz0,l(t),λl(t))p(s,X(0),i,dz0×{l})−Eh(XX(0),i(t),λi(t))|≤n∑l=1∫R2+|Eh(Xz0,l(t),λl(t))−Eh(XX(0),i(t),λi(t))|p(s,X(0),i,dz0×{l})≤2p(s,X(0),i,ˉDCR×S)+n∑l=1∫ˉDR|Eh(Xz0,l(t),λl(t))−Eh(XX(0),i(t),λi(t))|×p(s,X(0),i,dz0×{l}), | (2.21) |
where ˉDR={X∈R2+||X|≤R}, ˉDCR=R2+−ˉDR. According to the well-known Chebyshev inequality, the family of transition probabilities {p(t,X(0),i,dz0×{l}|t≥0)} is tight. That is to say, for any ε>0 there is a compact subset D=D(ε,X(0),i) of R2+ such that p(t,X(0),i,D×S)≥1−ε,∀t≥0, there is a positive number R sufficiently large for
p(s,X(0),i,ˉDCR×S)<ε34,∀s≥0. | (2.22) |
Notes that by Lemma 2.7, there is a T>0 such that
suph∈H|Eh(Xz0,l(t),λl(t))−Eh(XX0,i(t),λi(t))|<ε32,∀t≥T,(z0,l)∈ˉDR×S. | (2.23) |
Substituting (2.22) and (2.23) into (2.21), we get
|Eh(XX(0),i(t+s),λi(t+s))−Eh(XX(0),i(t),λi(t))|<ε3,∀t≥T,s>0. |
Because h is arbitrary, the inequality (2.20) must hold.
Proof of Theorem 2.4. First, we prove that there is a probability measure η(⋅×⋅)∈B such that for any (X(0),j)∈R2+×S, the transition probability p(t,X(0),j,⋅×⋅) of X((X(0),j),t) converges weakly to η(⋅×⋅). Based on proposition 2.5 in reference [35], the weak convergence of probability measures is a metric concept, namely, p(t,X(0),j,⋅×⋅) converges weakly to η(⋅×⋅)⇔ there is a metric d such that limt→+∞d(p(t,X(0),j,⋅×⋅),η(⋅×⋅))=0. Hence, we just need to show that for any (X(0),i)∈R2+×S,
limt→+∞d(p(t,X(0),j,⋅×⋅),η(⋅×⋅))=0. |
By using Lemma 2.8, {p(t,0,1,⋅×⋅)|t≥0} is Cauchy in the space B(R2+×S) with metric dH. Therefore, there is a unique η(⋅×⋅)∈B such that
limt→+∞dH(p(t,0,1,⋅×⋅),η(⋅×⋅))=0. |
By Lemma 2.7, we get
limt→+∞dH(p(t,X(0),j,⋅×⋅),η(⋅×⋅))≤limt→∞[dH(p(t,0,1,⋅×⋅),η(⋅×⋅))+dH(p(t,X(0),j,⋅×⋅),p(t,0,1,⋅×⋅))]=0. |
Namely, the distribution of (X(t),λ(t)) converges weakly to η. Due to the Kolmogorov-Chapman equation, we know that η is invariant. Applying the Corollary 3.43 in reference [36], we obtain that η is strong mixing. Therefore, η is ergodic by Theorem 3.2.6 and (3.3.2) in reference [36].
Corollary 2.1. Consider model (1.4), Assumptions 2.1–2.3 hold, then
(i) if ˉb1>0, ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗>0, then model (1.4) has a unique stationary distribution η(⋅×⋅).
(ii) if ˉb1>0, a11ˉb2+a21ˉb1<0, then x(t) has a unique ergodic stationary distribution, y(t) goes to extinction.
(iii) if ˉb1<0, a11ˉb2+a21ˉb1<0, then both x(t) and y(t) will go to extinction.
In this section, we will validate our theoretical results with the help of numerical simulation examples taking parameters. Our results show that the existence of stationary distribution has close relations with random disturbances. Let's consider the model (1.4) with S={1,2} to make it easier to understand. Therefore, according to the law of the Markov chain, system (1.4) can be regarded as a hybrid system which switches between the following two subsystems:
{dx(t)=x(t−)[r1(1)−a11x(t)−a12y(t)1+mx(t)+ny(t)]dt+σ1(1)x(t−)dB1(t)+∫Zx(t−)γ1(1,u)˜N(dt,du),dy(t)=y(t−)[−r2(1)−a22y(t)+a21x(t)1+mx(t)+ny(t)]dt+σ2(1)y(t−)dB2(t)+∫Zy(t−)γ2(1,u)˜N(dt,du). | (3.1) |
{dx(t)=x(t−)[r1(2)−a11x(t)−a12y(t)1+mx(t)+ny(t)]dt+σ1(2)x(t−)dB1(t)+∫Zx(t−)γ1(2,u)˜N(dt,du),dy(t)=y(t−)[−r2(2)−a22y(t)+a21x(t)1+mx(t)+ny(t)]dt+σ2(2)y(t−)dB2(t)+∫Zy(t−)γ2(2,u)˜N(dt,du). | (3.2) |
(i) Firstly, let us consider the effects about the distribution of Markov chain. We know the following facts: if the above two subsystems have a stationary distributions, then the hybrid system (1.4) still has a stationary distribution due to regime switching; if one of the two subsystems has a stationary distribution and the other does not, then the hybrid system (1.4) may have a stationary distribution, or not. To see the above more clearly, let's use several simulations to illustrate the impacts. Here, we only present the second case by letting the distribution of the Markov chain change(i.e., let π change).
Example 1. Choose a11=0.75,a12=0.25,a21=0.15,a22=0.95,Z=(0,+∞),m=0.15,n=0.2,λ(Z)=1. We have a11−a12mn−2a21=0.26≥0, a22−a21nm−2a12=0.25≥0.
In regime 1, choose r1(1)=0.08,r2(1)=0.20,σ1(1)=0.53,γ1(1)=0.15,σ2(1)=0.59,γ2(1)=0.10, thus b1(1)=−0.07<0,b2(1)=−0.38<0,a11b2(1)+a21b1(1)=−0.30<0. Therefore, according to Corollary 2.1, in subsystem (3.1), prey x(t) and predator y(t) go to extinction, see Figure 1.
In regime 2, choose r1(2)=0.45,r2(2)=0.01,σ1(2)=0.08,γ1(2)=0.09,σ2(2)=0.11,γ2(2)=0.11, therefore b1(2)=0.44>0,b2(2)=−0.02<0,a11b2(1)+a21b1(1)=0.05>0, but ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗>0 is difficult to verify. In subsystem (3.2), Figure 2 illustrates the situation that the model (1.4) has a unique stationary distribution η(⋅×⋅).
And then we're going to choose different π.
Case (a): We choose π=(0.050.95). Therefore ˉb1=0.05×(−0.07)+0.95×0.44=0.41>0; ˉb2=0.05×(−0.38)+0.95×(−0.02)=−0.04<0, a11ˉb2+a21ˉb1=0.75×(−0.04)+0.15×0.41=0.03>0, but ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗>0 is difficult to verify. Figure 3 illustrates that the model (1.4) has a unique stationary distribution η(⋅×⋅).
Case (b): We choose π=(0.30.7). Therefore ˉb1=0.3×(−0.07)+0.7×0.44=0.29>0; ˉb2=0.3×(−0.38)+0.7×(−0.02)=−0.13<0, and a11ˉb2+a21ˉb1=0.75×(−0.13)+0.15×0.29=−0.05<0. Thus according to Corollary 2.1, the prey x(t) has a unique ergodic stationary distribution, y(t) goes to extinction, see Figure 4.
Case (c): We choose π=(0.90.1). Therefore ˉb1=0.9×(−0.07)+0.1×0.44=−0.02<0; ˉb2=0.9×(−0.38)+0.1×(−0.02)=−0.34<0, and a11ˉb2+a21ˉb1=0.75×(−0.34)+0.15×(−0.02)=−0.26<0. Therefore, according to Corollary 2.1, prey x(t) and predator y(t) go to extinction, see Figure 5.
From the above examples, we can know the following facts: one subsystem has a stationary distribution while the other doesn't have, different things can happen when we choose different π. When π=(0.90.1), both prey x(t) and predator y(t) go to extinction; when π=(0.30.7), the prey x(t) has a unique ergodic stationary distribution, y(t) goes to extinction; when π=(0.050.95), the model (1.4) has a unique ergodic stationary distribution.
(ii) Next, we consider the effect of white noise on the population. For convenience, we will consider only subsystem (3.1). One can easily calculate the following facts
∂b1(1)∂σ21(1)<0,∂b2(1)∂σ22(1)<0. |
In short, subsystem (3.1) has a stationary distribution if for some σ21(1) and σ22(1), then the stationary distribution could disappear with the increase of σ21(1) or σ22(1). To see this more clearly, let's look at the following example.
Example 2. Consider the subsystem (3.1), the values of σ1(1) and σ2(1) are given in the table below, and the values of other parameters are the same with Example 1.
σ1(1) | σ2(1) | b1 | b2 | a11b2+a21b1 |
0.04 | 0.02 | 0.07 | -0.20 | -0.14 |
0.43 | 0.23 | -0.02 | -0.23 | -0.18 |
0.58 | 0.28 | -0.10 | -0.24 | -0.20 |
0.84 | 0.59 | -0.28 | -0.38 | -0.33 |
Therefore, in subsystem (3.1), according to Corollary 2.1, the prey x(t) has a unique ergodic stationary distribution with σ1(1)=0.04, x(t) is extinct at all other values. y(t) goes to extinction due to a11b2+a21b1<0, see Figure 6. We know that when σ1(1) and σ2(1) increase, the stationary distribution of prey x(t) could disappear, and the greater the intensity of random perturbation, the faster the species dies out.
(iii) Finally, consider the Lévy jumps. To keep things simple, we let γi(u)=εi(i=1,2). We have
ˉb1=∑j∈Sπj(r1(j)−12σ21(j))−(ε1−ln(1+ε1)),ˉb2=∑j∈Sπj(−r2(j)−12σ22(j))−(ε2−ln(1+ε2)), |
we know that εi−ln(1+εi)≥0, εi≥−1,i=1,2. Therefore, the effect of εi on the population is similar to that of σ2i, so it is omitted here.
After the above numerical simulation, there are similar examples in reality. In [37], the authors show that global warming has a profound bottom-up impact upon marine top-predators. Marine pollution via heavy metals, organochlorides, oil products and plastics is a recurrent threat to seabirds on a worldwide scale. All threats mentioned above cause substantial disturbance to seabird populations. Most seabirds feed on fish, and fish stocks are overexploited by industrial fishing, which can lead to large numbers of seabirds going hungry, all threats that cause significant disruption to bird populations. In [37], lesser sandeels Ammodytes marinus, which used to be the food-base of a vast seabird community around the British Isles, have been depleted due to the combined effects of overfishing and climate change. Seabirds, in particular kittiwakes Rissa tridactyla, now feed increasingly on snake pipefish. We know that not all seabirds are geographically malleable [38]. Seabirds may face extinction as some endemic diseases are trapped in restricted areas due to the effects of climate change. This is most likely the case for the Galˊapagos penguin Spheniscus mendiculus [39] and marblefinch bracketail fern [40]. Above examples show that under the influence of environmental disturbance caused by climate warming and Marine pollution, on the one hand, the population will die out under some certain conditions; on the other hand, under some certain conditions, the predator changes its foraging ecology so that it can survive, which indicates that the predator and the new prey will not die out.
For the predator-prey model with Beddington-DeAngelis functional response, it has important theoretical and practical significance in real life, and has received extensive attention. However, a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps has not been developed. In our work, our method easy to understand, but the sufficient conditions that make the conclusion hold may be a little strict, and the weakening of the conditions needs further study. Our main result is Theorem 2.4, which establishes sufficient conditions for the existence and uniqueness of an ergodic stationary distribution. Corollary 2.1 indicates that the existence of stationary distribution and extinction of model (1.4) depends on the signs of ˉb1, a11ˉb2+a21ˉb1 and ˉb2+⟨a21ˉx(t)1+mˉx(t)+nˉy(t)⟩∗. Our results show that the existence of stationary distribution has close relations with the disturbance of environment.
There are still some interesting questions that deserve further consideration. For example, maybe one can give the threshold between the weak persistence in the mean and extinction for predator in future, further weakening the conditions in Theorem 2.4. Another interesting question is to consider what happens if other parameters are perturbed by noises. It is also interesting to consider other population systems, such as, food chain models, competition model and so on.
We would like to thank the the Fundamental Research Funds for the Central Universities (Grant number 3122021073).
All authors declare no conflicts of interest in this paper.
[1] |
M. Liu, C. X. Du, M. L. Deng, Persistence and extinction of a modified Leslie-Gower Holling-type II stochastic predator-prey model with impulsive toxicant input in polluted environments, Nonlinear Anal.-Hybri., 27 (2018), 177–190. http://doi.org/10.1016/j.nahs.2017.08.001 doi: 10.1016/j.nahs.2017.08.001
![]() |
[2] |
Z. J. Wang, M. L. Deng, M. Liu, Stationary distribution of a stochastic ratio-dependent predator-prey system with regime-switching, Chaos Soliton. Fract., 142 (2021), 110462. https://doi.org/10.1016/j.chaos.2020.110462 doi: 10.1016/j.chaos.2020.110462
![]() |
[3] | R. J. Swift, A stochastic predator-prey model, Bull. Irish Math. Soc., 48 (2002), 57–63. |
[4] |
Q. Liu, D. Q. Jiang, T. Hayat, Dynamics of stochastic predator-prey models with distributed delay and stage structure for prey, Int. J. Biomath., 14 (2021), 2150020. https://doi.org/10.1142/S1793524521500200 doi: 10.1142/S1793524521500200
![]() |
[5] | S. Q. Zhang, T. H. Zhang, S. L. Yuan, Dynamics of a stochastic predator-prey model with habitat complexity and prey aggregation, Ecol. Complexity. 45 (2021), 100889. https://doi.org/10.1016/j.ecocom.2020.100889 |
[6] |
L. L. Jia, Analysis for a delayed three-species predator-prey model with feedback controls and prey diffusion, J. Math., 2020 (2020), 5703859. https://doi.org/10.1155/2020/57038599 doi: 10.1155/2020/57038599
![]() |
[7] |
J. R. Beddington, Mutual interference between parasites or predators and its effect on searching efficiency, J. Anim. Ecol., 44 (1975), 331–341. https://doi.org/10.2307/3866 doi: 10.2307/3866
![]() |
[8] |
D. L. DeAngelis, R. A. Goldsten, R. V. O'Neill, A model for trophic interaction, Ecology, 56 (1975), 881–892. https://doi.org/10.2307/1936298 doi: 10.2307/1936298
![]() |
[9] |
M. Liu, P. S. Mandal, Dynamical behavior of a one-prey two-predator model with random perturbations, Commun. Nonlinear Sci., 28 (2015), 123–137. https://doi.org/10.1016/j.cnsns.2015.04.010 doi: 10.1016/j.cnsns.2015.04.010
![]() |
[10] |
N. H. Du, V. H. Sam, Dynamics of a stochastic Lotka-Volterra model perturbed by white noise, J. Math. Anal. Appl., 324 (2006), 82–97. https://doi.org/10.1016/j.jmaa.2005.11.064 doi: 10.1016/j.jmaa.2005.11.064
![]() |
[11] |
D. Q. Jiang, N. Z. Shi, X. Y. Li, Global stability and stochastic permanence of a nonautonomous logistic equation with random perturbation, J. Math. Anal. Appl., 340 (2008), 588–597. https://doi.org/10.1016/j.jmaa.2007.08.014 doi: 10.1016/j.jmaa.2007.08.014
![]() |
[12] |
X. R. Mao, G. Marion, E. Renshaw, Environmental Brownian noise suppresses explosions in population dynamics, Stoch. Proc. Appl., 97 (2002), 95–110. https://doi.org/10.1016/S0304-4149(01)00126-0 doi: 10.1016/S0304-4149(01)00126-0
![]() |
[13] | X. R. Mao, C. G. Yuan, Stochastic differential equations with Markovian switching, London: Imperial College Press, 2006. https://doi.org/10.1142/p473 |
[14] |
H. Wang, M. Liu, Stationary distribution of a stochastic hybrid phytoplankton-zooplankton model with toxin-producing phytoplankton, Appl. Math. Lett., 101 (2020), 106077. https://doi.org/10.1016/j.aml.2019.106077 doi: 10.1016/j.aml.2019.106077
![]() |
[15] |
D. G. Li, M. Liu, Invariant measure of a stochastic food-limited population model with regime switching, Math. Comput. Simulat., 178 (2020), 16–26. https://doi.org/10.1016/j.matcom.2020.06.003 doi: 10.1016/j.matcom.2020.06.003
![]() |
[16] |
C. Jeffries, Stability of predation ecosystem models, Ecology, 57 (1976), 1321–1325. https://doi.org/10.2307/1935058 doi: 10.2307/1935058
![]() |
[17] |
J. H. Bao, J. H. Shao, Permanence and extinction of regime-switching predator-prey models, SIAM J. Math. Anal., 48 (2016), 725–739. https://doi.org/10.1137/15M1024512 doi: 10.1137/15M1024512
![]() |
[18] |
C. Zhu, G. Yin, On competitive Lotka-Volterra model in random environments, J. Math. Anal. Appl., 357 (2009), 154–170. https://doi.org/10.1016/j.jmaa.2009.03.066 doi: 10.1016/j.jmaa.2009.03.066
![]() |
[19] |
M. Liu, Dynamics of a stochastic regime-switching predator-prey model with modified Leslie-Gower Holling-type II schemes and prey harvesting, Nonlinear Dyn., 96 (2019), 417–442. https://doi.org/10.1007/s11071-019-04797-x doi: 10.1007/s11071-019-04797-x
![]() |
[20] |
J. H. Bao, X. R. Mao, G. G. Yin, C. G. Yuan, Competitive Lotka-Volterra population dynamics with jumps, Nonlinear Anal.-Theor, 74 (2011), 6601–6616. https://doi.org/10.1016/j.na.2011.06.043 doi: 10.1016/j.na.2011.06.043
![]() |
[21] |
J. H. Bao, C. G. Yuan, Stochastic population dynamics driven by Lévy noise, J. Math. Anal. Appl., 391 (2012), 363–375. https://doi.org/10.1016/j.jmaa.2012.02.043 doi: 10.1016/j.jmaa.2012.02.043
![]() |
[22] | Y. C. Zang, J. P. Li, A dynamics of a stochastic predator-prey system with Beddington-DeAngelis functional response and Lévy jumps, Acta Math. Appl. Sin., 38 (2015), 340–349. |
[23] |
H. Qiu, M. Liu, K. Wang, Y. Wang, Dynamics of a stochastic predator-prey system with Beddington-DeAngelis functional response, Appl. Math. Comput., 219 (2012), 2303–2312. https://doi.org/10.1016/j.amc.2012.08.077 doi: 10.1016/j.amc.2012.08.077
![]() |
[24] | V. Y. Krasin, A. V. Melnikov, On comparison theorem and its applications to finance, In: Optimality and risk modern trends in mathematical Finance, Berlin: Springer, 2009. https://doi.org/10.1007/978-3-642-02608-98 |
[25] |
J. H. Bao, C. G. Yuan, Comparison theorem for stochastic differential delay equations with jumps, Acta Appl. Math., 116 (2011), 119. https://doi.org/10.1007/s10440-011-9633-7 doi: 10.1007/s10440-011-9633-7
![]() |
[26] |
X. L. Zou, K. Wang, Optimal harvesting for a stochastic regime-switching logistic diffusion system with jumps, Nonlinear Anal. Hybri., 13 (2014), 32–44. https://doi.org/10.1016/j.nahs.2014.01.001 doi: 10.1016/j.nahs.2014.01.001
![]() |
[27] |
M. Liu, Y. Zhu, Stationary distribution and ergodicity of a stochastic hybrid competition model with Lévy jumps, Nonlinear Anal. Hybri., 30 (2018), 225–239. https://doi.org/10.1016/j.nahs.2018.05.002 doi: 10.1016/j.nahs.2018.05.002
![]() |
[28] |
M. Liu, C. z. Bai, Dynamics of a stochastic one-prey two-predator model with Lévy jumps, Appl. Math. Comput., 284 (2016), 308–321. https://doi.org/10.1016/j.amc.2016.02.033 doi: 10.1016/j.amc.2016.02.033
![]() |
[29] |
S. Li, X. N. Zhang, Dynamics of a stochastic non-autonomous predator-prey system with Beddington-DeAngelis functional response, Adv. Differ. Equ., 2013 (2013), 19. https://doi.org/10.1186/1687-1847-2013-19 doi: 10.1186/1687-1847-2013-19
![]() |
[30] |
R. Lipster, A strong law of large numbers for local martingales, Stochastics, 3 (1980), 217–228. https://doi.org/10.1080/17442508008833146 doi: 10.1080/17442508008833146
![]() |
[31] |
J. L. Lv, K. Wang, M. Liu, Dynamical properties of a stochastic two-species Schoener's competitive model, Int. J. Biomath., 5 (2012), 1250035. https://doi.org/10.1142/S1793524511001751 doi: 10.1142/S1793524511001751
![]() |
[32] | I. Barbǎlat, Systemes d'equations differential d'oscillations nonlinearies (in Romanian), Revue des Math. Pures et Appl., 4 (1959), 267–270. |
[33] | X. R. Mao, Stochastic differential equations and applications, 2 Eds., England: International Publishers in Science and Technology, 2007. |
[34] | D. Applebaum, Lévy processes and stochastic calculus, 2 Eds., Cambridge: Cambridge Unversity Press, 2009. https://doi.org/10.1017/CBO9780511809781 |
[35] |
G. B. Kallianpur, Stochastic differential equations and diffusion processes, Technometrics, 25 (1983), 208. https://doi.org/10.1080/00401706.1983.10487861 doi: 10.1080/00401706.1983.10487861
![]() |
[36] | G. Da Prato, J. Zabczyk, Ergodicity for infinite dimensional systems, Cambridge: Cambridge University Press, 1996. https://doi.org/10.1017/CBO9780511662829 |
[37] |
E. S. Poloczanska, R. M. Cook, G. D. Ruxton, P. J. Wright, Fishing vs. natural recruitment variation in sandeels as a cause of seabird breeding failure at Shetland: A modelling approach, Ices J. Mar. Sci., 61 (2004), 788–797. https://doi.org/10.1016/j.icesjms.2004.03.030 doi: 10.1016/j.icesjms.2004.03.030
![]() |
[38] |
D. Gremillet, T. Boulinier, Spatial ecology and conservation of seabirds facing global climate change: a review, Mar. Ecol. Prog. Ser., 391 (2009), 121–137. https://doi.org/10.3354/meps08212 doi: 10.3354/meps08212
![]() |
[39] |
F. H. Vargas, R. C. Lacy, P. J. Johnson, A. Steinfurth, R. J. M. Crawford, P. D. Boersma, et al., Modelling the effect of El Nio on the persistence of small populations: The Galapagos penguin as a case study, Biol. Conserv., 137 (2007), 138–148. https://doi.org/10.1016/j.biocon.2007.02.005 doi: 10.1016/j.biocon.2007.02.005
![]() |
[40] |
B. H. Becker, M. Z. Peery, S. R. Beissinger, Ocean climate and prey availability affect the trophic level and reproductive success of the marbled murrelet, an endangered seabird, Mar. Ecol. Prog. Ser., 329 (2007), 267–279. https://doi.org/10.3354/meps329267 doi: 10.3354/meps329267
![]() |
1. | Jawdat Alebraheem, Yahya Abu-Hassan, A novel mechanism measurement of predator interference in predator–prey models, 2023, 86, 0303-6812, 10.1007/s00285-023-01914-8 |
σ1(1) | σ2(1) | b1 | b2 | a11b2+a21b1 |
0.04 | 0.02 | 0.07 | -0.20 | -0.14 |
0.43 | 0.23 | -0.02 | -0.23 | -0.18 |
0.58 | 0.28 | -0.10 | -0.24 | -0.20 |
0.84 | 0.59 | -0.28 | -0.38 | -0.33 |