Conditions | Species y1 | Species y2 |
ˉη1<0 | Extinction | Extinction |
ˉη1>0 and Δ2<0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | Extinction |
Δ2>0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | limt→∞1t∫t0y2(s)ds=Δ2A11A22 |
This work is concerned with a stochastic predator-prey system with S-type distributed time delays, regime switching and Lévy jumps. By use of the stochastic differential comparison theory and some inequality techniques, we study the extinction and persistence in the mean for each species, asymptotic stability in distribution and the optimal harvesting effort of the model. Then we present some simulation examples to illustrate the theoretical results and explore the effects of regime switching, distributed time delays and Lévy jumps on the dynamical behaviors, respectively.
Citation: Yuanfu Shao. Dynamics and optimal harvesting of a stochastic predator-prey system with regime switching, S-type distributed time delays and Lévy jumps[J]. AIMS Mathematics, 2022, 7(3): 4068-4093. doi: 10.3934/math.2022225
[1] | Hong Qiu, Yanzhang Huo, Tianhui Ma . Dynamical analysis of a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps. AIMS Mathematics, 2022, 7(8): 14492-14512. doi: 10.3934/math.2022799 |
[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] | 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 |
[4] | 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 |
[5] | 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 |
[6] | Yassine Sabbar, Aeshah A. Raezah, Mohammed Moumni . Enhancing epidemic modeling: exploring heavy-tailed dynamics with the generalized tempered stable distribution. AIMS Mathematics, 2024, 9(10): 29496-29528. doi: 10.3934/math.20241429 |
[7] | Jingjing Yang, Jianqiu Lu . Stabilization in distribution of hybrid stochastic differential delay equations with Lévy noise by discrete-time state feedback controls. AIMS Mathematics, 2025, 10(2): 3457-3483. doi: 10.3934/math.2025160 |
[8] | Lin Xu, Linlin Wang, Hao Wang, Liming Zhang . Optimal investment game for two regulated players with regime switching. AIMS Mathematics, 2024, 9(12): 34674-34704. doi: 10.3934/math.20241651 |
[9] | Dennis Llemit, Jose Maria Escaner IV . Value functions in a regime switching jump diffusion with delay market model. AIMS Mathematics, 2021, 6(10): 11595-11609. doi: 10.3934/math.2021673 |
[10] | Hui Sun, Zhongyang Sun, Ya Huang . Equilibrium investment and risk control for an insurer with non-Markovian regime-switching and no-shorting constraints. AIMS Mathematics, 2020, 5(6): 6996-7013. doi: 10.3934/math.2020449 |
This work is concerned with a stochastic predator-prey system with S-type distributed time delays, regime switching and Lévy jumps. By use of the stochastic differential comparison theory and some inequality techniques, we study the extinction and persistence in the mean for each species, asymptotic stability in distribution and the optimal harvesting effort of the model. Then we present some simulation examples to illustrate the theoretical results and explore the effects of regime switching, distributed time delays and Lévy jumps on the dynamical behaviors, respectively.
The natural world is full of environmental perturbations almost everywhere, which play an important role in the ecological system [1,2,3,4,5]. As May [6] said that the growth rates in ecological systems should be stochastic due to the influences of random noises, so the solution would fluctuate around some average values, not being a steady positive point. Therefore, it is more rational by considering a stochastic environmental perturbation (white noise) to better investigate the real population systems. In mathematical modelling, the white noise has been extensively introduced to reveal the stochastic population dynamics, see e.g. [3,4,5,7,8] and references cited therein.
In real world, population systems may inevitably suffer some abrupt changes. It is well known that the growth rates of some species are affected by seasonal factors, which are much different in summer from those in winter. These phenomena can be described by a continuous-time Markovian process with a finite state space, i.e. colorful noise in mathematical modelling. The colorful noise may take several values and switch among different regimes of environment, which is memoryless, and the waiting time for the next switching follows an exponential distribution. The effect of colorful noise on population dynamics has attracted many researchers [9,10,11,12,13,14]. For example, by using a Markovian switching process to model the colorful noise in environment, Liu, He and Yu [14] proposed the following stochastic system with harvesting and regime-switching:
{dx1(t)=x1(t)(r1(α(t))−h1−a11x1(t)−a12∫0−τ1x2(t+s)dμ1(s))dt+σ1(α(t))x1(t)dB1(t),dx2(t)=x2(t)(r2(α(t))−h2+a21∫0−τ2x1(t+s)dμ2(s)−a22x2(t))dt+σ2(α(t))x2(t)dB2(t), | (1.1) |
where r1(⋅)>0 is the growth rate of prey, r2(⋅)<0 is the death rate of predator, aii>0 is the intra-specific competition rate, a12>0 is the capture rate and a21>0 is the conversion rate of food; hi is the harvesting constant; B1(t) and B2(t) are standard independent Brownian motions defined on a complete probability space (Ω,F,{Ft}t≥0,P), where {Ft}t≥0 is a filtration. It is right continuous and F0 contains all P-null sets; σ2i(⋅) is the intensity of stochastic noise, i=1,2. The regime of switching α(t) is a Markovian chain in a finite state space S={1,2,…,N}. The generator of α(t) is defined as Γ=(γij)N×N with
P{α(t+△t=j|α(t)=i)}={γij△t+o(△t),i≠j,1+γii△t+o(△t),i=j, |
where △t>0,γij is the transition rate from the ith stage to the jth stage and γij≥0 if i≠j while γii=−∑i≠jγij,i,j∈S.
It is often assumed that every sample of α(t) is a right continuous step function and irreducible with a finite simple jumps in any finite subinterval of R+=[0,∞). It obeys a unique stationary distribution π=(π1,π2,…,πN) satisfying πΓ=0 and ∑Nk=1πk=1,πk>0,k∈S. The switching mechanism of the hybrid system is referred to [9].
In the study of ecological dynamics, the current growth of populations is usually influenced by its past history, that is, time delay is often inevitable in the natural ecosystems [15]. S-type distributed time delay is such a distributed delay that the integral is Lebesgue-Stieltjes. Just as the authors [16] said, "systems with discrete time delays and those with continuously distributed time delays do not contain each other. However, systems with S-type distributed time delays contain both." So it is interesting to consider the impact of S-type distributed delay on the ecological dynamics. Models with S-type distributed time delays have been studied by many authors [17,18]. On the other hand, earthquake, harvesting and epidemics often happen in natural world. These sudden environmental perturbations are so strong and can change the population size in a very short time, which can not be described by white noise [19,20]. Many experiments show that, due to the influence of environmental disturbance, the distribution of many species exhibits a scale-free characteristic, and hence the biologists introduce a non-Gaussian Lévy jump to characterize it. Models with Lévy jump are studied by many researchers and many nice results have been obtained [21,22,23].
Considering the effects of S-type time delays and Lévy jumps on system (1.1), we establish the following stochastic model
{dx1(t)=x1(t−)(r1(α(t−))−h1−a11x1(t−)−∫0−τ11x1(t+θ)dμ11(θ)−a12x2(t−)−∫0−τ12x2(t+θ)dμ12(θ))dt+σ1(α(t−))x1(t−)dB1(t)+∫Zγ1(u,α(t−))x1(t−)˜N(dt,du),dx2(t)=x2(t−)(r2(α(t−))−h2+a21x1(t−)+∫0−τ21x1(t+θ)dμ21(θ)−a22x2(t−)−∫0−τ22x2(t+θ)dμ22(θ))dt+σ2(α(t−))x2(t−)dB2(t)+∫Zγ2(u,α(t−))x2(t−)˜N(dt,du), | (1.2) |
with initial value
x(θ)=(x1(θ),x2(θ))T=(ϕ1(θ),ϕ2(θ))T=ϕ(θ)∈C([−τ,0];R2+),α(θ)=ς, |
where xi(t−) stands for the left limit of xi(t); ∫0−τijxi(t+θ)dμij(θ) is Lebesgue-Stieltjes integral; τij>0 is time delay; μij(θ) is bounded, nondecreasing variation function defined on [−τ,0] with τ=maxi,j=1,2{τij},ς∈S,i,j=1,2; R2+={x=(x1,x2)∈R2,xi>0,i=1,2} with |x(t)|=√∑2i=1x2i; N is a Poisson counting measure, ˜N(dt,du)=N(dt,du)−λ(du) is the component of N, where λ is the characteristic measure on a measurable subset Z⊂R+=[0,∞) such that λ(Z)<∞. The Markov chain, Brownian motion and Lévy jumps are mutually independent.
Our main goal of this paper is as follows. First, since the study of dynamical behaviors of predator-prey system is an important topic [7,24], we establish some sufficient conditions assuring the extinction, persistence in the mean for all species of system (1.2).
Second, in the study of long-term behaviors of species, the existence of a unique probability measure plays an important role in stochastic models with Lévy jumps (see [25,26,27]). Hence, it is very interesting to analyze the asymptotic stability in distribution of (1.2).
Third, in mathematical biology, it is valuable to keep the species persistent to maintain the biological balance. Consequently, the optimal harvesting strategy of renewable resources becomes more and more important [8,27,28,29]. By ergodic method, we will study the optimal harvesting strategy of system (1.2).
The rest of this paper is organized as follows. Section 2 begins with some definitions, important lemmas and notations. Section 3 is devoted to the extinction and persistence in the mean for species. Section 4 and Section 5 focus on the asymptotic stability in distribution and the optimal harvesting strategy of (1.2), respectively. Section 6 gives some numerical simulations to verify the main results. Finally, Section 7 presents a brief discussion to conclude this paper.
To begin with this section, we introduce some notations of the Itô's integral for a stochastic differential equation with Markovian switching and Lévy jumps [19,23]. Let
dx(t)=f(x(t−),t−,α(t−))dt+g(x(t−),t−,α(t−))dB(t)+∫Zh(x(t−),t−,α(t−),μ)˜N(dt,dμ), |
where fandg:R2×R+×S→R2,h:R2×R+×S×Z→R2 are measurable functions. Let V∈C2,1(R2×R+×S,R2). Define the operator L as follows:
LV(x,t,k)=Vt(x,t,k)+Vx(x,t,k)f(x,t,k)+12trace[gT(x,t,k)Vxx(x,t,k)g(x,t,k)]+∫Z{V(x+h(x,t,k,u),t,k)−V(x,t,k)−Vx(x,t,k)h(x,t,k,u)}λ(du)+N∑j=1γkjV(x,t,j), |
where Vt(x,t,k)=∂V(x,t,k)∂t,Vx(x,t,k)=(∂V(x,t,k)∂x1,∂V(x,t,k)∂x2),Vxx(x,t,k)=(∂2V(x,t,k)∂xi∂xj)2×2,i,j=1,2. The generalized Itô's formula with jumps (see, for example[19,23]) is defined as
dV(x,t,k)=LV(x,t,k)dt+Vx(x,t,k)g(x,t,k)dB(t)+∫Z{V(x+h(x,t,k,u),t,k)−V(x,t,k)}˜N(dt,du). |
Now we give the definitions of extinction and persistence in the mean of each species, and an important comparison theorem.
Definition 2.1 ([27]). Let x(t)=(x1(t),x2(t))T∈R2+ be a solution of system (1.2). Then,
(a) the population xi(t) is said to be extinct if limt→∞xi(t)=0,i=1,2;
(b) the population xi(t) is said to be persistent in the mean if limt→∞1t∫t0xi(s)ds=K a.s., where K is a positive constant, i=1,2.
Lemma 2.1 ([4]). Suppose that Z(t)∈C[Ω×[0,+∞),R+] and limt→∞F(t)/t=0,a.s.
(a) If there exist two positive constants T and λ0 such that, for all t>T,
lnZ(t)≤λt−λ0∫t0Z(s)ds+F(t),a.s., |
then
{lim supt→∞1t∫t0Z(s)ds≤λ/λ0,a.s.,ifλ≥0,limt→∞Z(t)=0,a.s.,ifλ<0. |
(b) If there exist three positive constants T,λ0 and λ such that, for all t>T,
lnZ(t)≥λt−λ0∫t0Z(s)ds+F(t),a.s, |
then
lim inft→∞1t∫t0Z(s)ds≥λ/λ0,a.s. |
Further, for the need of our discussion, we give some technical assumptions.
Assumption 1. For any α∈S and i=1,2, we assume γi(α,u)>−1 and
∫Z{γi(α,u)−ln(1+γi(α,u))}λ(du)≤K, |
∫Zmax{|γi(α,u)|2,[ln(1+γi(α,u))]2}λ(du)≤K, |
where K>0 denotes a positive and finite constant unless otherwise stated, which may be different in different places.
Assumption 2. For any p>0, we assume
∫Z{|1+γi(α,u)|p−1−pγi(α,u)}λ(du)≤K(p). |
Assumptions 1 and 2 imply that the intensity of Lévy noise cannot be too strong, otherwise, the solution of system (1.2) may explode in some finite time [23]. About the existence of nontrivial positive solutions of (1.2), we have the following lemma.
Lemma 2.2. Let Assumptions 1 and 2 hold. For any given initial value x(θ)∈C([−τ,0];R2+),α(θ)=ς, system (1.2) has a unique solutionx(t) on t∈[−τ,∞), and the solution remains in R2+ with probability one. Moreover, for any p>0, there is a K(p)>0such that lim supt→∞E|x(t)|p≤K(p).
Proof. The proof of the existence of solutions is straightforward, one may refer to [14,23]. As to the proof of the boundedness of expectation, one can get it by using the generalized Itô's formular with jumps to function et∑2i=1xpi(t). The process is similar to reference [14] and is omitted.
For simplicity, we introduce some notations to end this section.
ξi(α(t))=ri(α(t))−σ2i(α(t))2−∫Z[γi(u,α(t))−ln(1+γi(u,α(t)))]λ(du),¯ξi=N∑k=1πkξi(k), |
ηi(α(t))=ξi(α(t))−hi,¯ηi=N∑k=1πkηi(k),Aij=aij+∫0−τijdμij(θ),i,j=1,2, |
Δ=|A11A12−A21A22|,Δ1=|¯η1A12¯η2A22|,Δ2=|A11¯η1−A21¯η2|. |
In this section, we study the long-term behaviors of (1.2). Consider the following system,
dW(t)=W(t−)(r(α(t−))−aW(t−)−∫0−τW(t−+θ)dμ(θ))dt+σ(α(t−))W(t−)dB(t)+∫Zγ(α(t−),u)W(t−)˜N(dt,du). | (3.1) |
Lemma 3.1. For system (3.1), the following statements hold.
(i) If ∑Nk=1πkξ(k)<0, then limt→∞W(t)=0.
(ii) If ∑Nk=1πkξ(k)>0, thenlimt→∞1t∫t0W(s)ds=∑Nk=1πkξ(k)a+∫0−τdμ(θ),
where ξ(k)=r(k)−σ2(k)2−∫Z[γ(k,u)−ln(1+γ(k,u))]λ(du).
Proof. Using the Itô's formula to lnW(t) and computing the stochastic differential along the solution of (3.1), we have
dlnW(t)=(r(α(t))−aW(t)−∫0−τW(t+θ)dμ(θ)−∫Z[γ(u,α(t))−ln(1+γ(α(t)))]λ(du)−12σ2(α(t)))dt+σ(α(t))dB(t)+∫Zln(1+γ(u,α(t)))˜N(dt,du). | (3.2) |
Integrating both sides of (3.2) from 0 to t, and divided by t from both sides, then
1tlnW(t)W(0)=1t∫t0ξ(α(s))ds−a1t∫t0W(s)ds−1t∫0−τ[∫0θ+∫t0+∫t+θt]W(s)dsdμ(θ)+1t∫t0σ(α(s))dB(s)+1t∫t0∫Zln(1+γ(u,α(s)))˜N(ds,du). | (3.3) |
Since
limt→∞1t∫0−τ∫t+θ0W(s)dsdμ(θ)=limt→∞∫0−τdμ(θ)×1t∫t0W(s)ds, |
it follows that
limt→∞1t∫0−τ∫t+θtW(s)dsdμ(θ)=limt→∞1t∫0−τ∫t+θ0W(s)dsdμ(θ)−limt→∞1t∫0−τdμ(θ)∫t0W(s)ds=0. |
Then
limt→∞1t∫0−τ[∫0θW(s)ds−∫tt+θW(s)ds]dμ(θ)=limt→∞1t∫0−τ[∫0θψ(s)ds−∫tt+θW(s)ds]dμ(θ)=0. |
Therefore (3.3) leads to (3.4) as follows.
1tlnW(t)W(0)=1t∫t0ξ(α(s))ds−(a+∫0−τdμ(θ))1t∫t0W(s)ds+1t∫t0σ(α(s))dB(s)+1t∫t0∫Zln(1+γ(u,α(s)))˜N(ds,du). | (3.4) |
On the other hand, by the ergodicity of Markovian chain, we have
limt→∞1t∫t0ξ(α(s))ds=N∑k=1πkξ(k). |
Now we give the proof of (ⅰ) and (ⅱ).
(ⅰ) By comparison method, using Lemma 2.1, we can easily derive from (3.4) that
lim supt→∞1t∫t0W(s)ds≤N∑k=1πkξ(k)a+∫0−τdμ(θ). |
If ∑Nk=1πkξ(k)<0, it is clear that limt→∞W(t)=0.
(ⅱ) If ∑Nk=1πkξ(k)>0, by Lemma 2.1 again, we have
lim inft→∞1t∫t0W(s)ds≥N∑k=1πkξ(k)a+∫0−τdμ(θ). |
Therefore,
limt→∞1t∫t0W(s)ds=N∑k=1πkξ(k)a+∫0−τdμ(θ). |
The proof is completed.
Next, we consider the following comparison system of (1.2):
{dy1(t)=y1(t)(r1(α(t))−h1−a11y1(t)−∫0−τ11y1(t+s)dμ11(s))dt+σ1(α(t))y1(t)dB1(t)+∫Zγ1(u,α(t))y1(t)˜N(dt,du),dy2(t)=y2(t)(r2(α(t))−h2+a21y1(t)+∫0−τ21y1(t+s)dμ21(s)−a22y2(t)−∫0−τ22y2(t+s)dμ22(s))dt+σ2(α(t))y2(t)dB2(t)+∫Zγ2(u,α(t))y2(t)˜N(dt,du). | (3.5) |
By Lemma 3.1, we derive from (3.5) that
limt→∞1t∫t0y1(s)ds=N∑k=1πkη1(k)a11+∫0−τ11dμ11(θ)≜ˉη1A11. |
Consider the following comparison system of the second equation of (3.5),
dˆy2(t)=ˆy2(t−)(r2(α(t))−h2+a21y1(t−)+∫0−τ21y1(t+s)dμ21(s)−a22ˆy2(t−))dt+σ2(α(t−))ˆy2(t−)dB(t)+∫Zγ2(u,α(t−))ˆy2(t−)˜N(dt,du). |
Similar with the previous reasoning, using Lemma 3.1 again, we can derive that
limt→∞1t∫t0ˆy2(s)ds=A11ˉη2+A21ˉη1A11A22≜Δ2A11A22. |
By the comparison theory [1,9], it is clear that x1≤y1 and y2≤ˆy2. Consequently, we have the following results.
Lemma 3.2. For system (3.5), the following statements of Table 1 hold.
Conditions | Species y1 | Species y2 |
ˉη1<0 | Extinction | Extinction |
ˉη1>0 and Δ2<0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | Extinction |
Δ2>0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | limt→∞1t∫t0y2(s)ds=Δ2A11A22 |
Now we give the main result on the extinction and persistence in the mean of the species of system (1.2).
Theorem 3.1. For system (1.2), the following statements hold (see the Table 2).
Cases | Conditions | Species x1 | Species x2 |
ⅰ | ˉη1<0 | Extinction | Extinction |
ⅱ | ˉη1>0 and Δ2<0 | limt→∞1t∫t0x1(s)ds=ˉη1A11 | Extinction |
ⅲ | Δ2>0 | limt→∞1t∫t0x1(s)ds=Δ1Δ | limt→∞1t∫t0x1(s)ds=Δ2Δ |
Proof. For system (1.2), by using the Itô's formula to lnxi(t),i=1,2, we have
dlnx1(t)=(r1(α(t))−h1−a11x1(t)−∫0−τ11x1(t+θ)dμ11(θ)−a12x2(t)−∫0−τ12x2(t+θ)dμ12(θ)−12σ21(α(t))−∫Z[γ1(u,α(t))−ln(1+γ1(α(t)))]λ(du))dt+σ1(α(t))dB1(t)+∫Zln(1+γ1(u,α(t)))˜N(dt,du), | (3.6) |
and
dlnx2(t)=(r2(α(t))−h2+a21x1(t)+∫0−τ21x1(t+θ)dμ21(θ)−a22x2(t)−∫0−τ22x2(t+θ)dμ22(θ)−12σ22(α(t))−∫Z[γ2(u,α(t))−ln(1+γ2(α(t)))]λ(du))dt+σ2(α(t))dB2(t)+∫Zln(1+γ2(u,α(t)))˜N(dt,du). | (3.7) |
Similar with the proof of Lemma 3.1, we have
limt→∞1t∫0−τ11[∫0θ+∫t0+∫t+θt]x1(s)dsdμ11(θ)=limt→∞1t∫0−τ11[∫0θx1(s)ds−∫tt+θx1(s)ds]dμ11(θ)+limt→∞1t∫0−τ11∫t0x1(s)dsdμ11(θ)=limt→∞1t∫0−τ11dμ11(θ)∫t0x1(s)ds, | (3.8) |
and
limt→∞1t∫0−τ12[∫0θ+∫t0+∫t+θt]x2(s)dsdμ12(θ)=limt→∞1t∫0−τ12[∫0θx2(s)ds−∫tt+θx2(s)ds]dμ12(θ)+limt→∞1t∫0−τ12∫t0x2(s)dsdμ12(θ)=limt→∞1t∫0−τ12dμ12(θ)∫t0x2(s)ds. | (3.9) |
Integrating both sides of (3.6) from 0 to t, and combining (3.8) and (3.9), then
1tlnx1(t)x1(0)=1t∫t0η1(α(s))ds−a111t∫t0x1(s)ds−1t∫0−τ11[∫0θ+∫t0+∫t+θt]x1(s)dsdμ11(θ)−a121t∫t0x2(s)ds−1t∫0−τ12[∫0θ+∫t0+∫t+θt]x2(s)dsdμ12(θ)+1t∫t0σ1(α(s))dB1(s)+1t∫t0∫Zln(1+γ1(u,α(s)))˜N(ds,du).=1t∫t0η1(α(s))ds−a11t∫t0x1(s)ds−1t∫0−τ11dμ11(θ)∫t0x1(s)ds−a12t∫t0x2(s)ds−1t∫0−τ12dμ12(θ)∫t0x2(s)ds+1t∫t0σ1(α(s))dB1(s)+1t∫t0∫Zln(1+γ1(u,α(s)))˜N(ds,du). | (3.10) |
By the same argumentation, we have
1tlnx2(t)x2(0)=1t∫t0η2(α(s))ds+a21t∫t0x1(s)ds+1t∫0−τ21dμ21(θ)∫t0x1(s)ds−a22t∫t0x2(s)ds−1t∫0−τ22dμ22(θ)∫t0x2(s)ds+1t∫t0σ2(α(s))dB2(s)+1t∫t0∫Zln(1+γ2(u,α(s)))˜N(ds,du). | (3.11) |
By the ergodicity of Markovian chain, then
limt→∞1t∫t0ηi(α(s))ds=N∑k=1πkηi(k),i=1,2. |
Now we prove the conclusion of Theorem 3.1.
(ⅰ) If ˉη1<0, then by Lemma 3.2, we have limt→∞x1(t)=0. Since ˉη2<0, it follows from (3.11) that limt→∞x2(t)=0. Hence limt→∞xi(t)=0,i=1,2.
(ⅱ) If ˉη1>0, then we derive from Lemma 3.2 that lim supt→∞1t∫t0x1(s)ds≤ˉη1A11. Substituting it into (3.11) and using Lemma 2.1 gives
lim supt→∞1t∫t0x2(s)ds≤¯η2+A21ˉη1A11A22=A11ˉη2+A21ˉη1A11A22=Δ2A11A22. |
By the condition Δ2<0, then limt→∞x2(t)=0.
From (3.11) again, by the comparison theorem [4,14,27], we have lim inft→∞1t∫t0x1(s)ds≥ˉη1A11. Therefore,
limt→∞1t∫t0x1(s)ds=ˉη1A11. |
(ⅲ) If Δ2>0, it is clear that ˉη1>0. We compute (3.10)×A21+(3.11)×A11, then
A211tlnx1(t)x1(0)+A111tlnx2(t)x2(0)=A21t∫t0(η1(α(s))ds+A11t∫t0(η2(α(s))ds−(A12A21+A11A22)t∫t0x2(s)ds+1t[A21M1(t)+A11M2(t)], |
where Mi(t)=∫t0σi(α(s))dBi(s)+∫t0∫Zln(1+γi(u,α(s)))˜N(ds,du),i=1,2. By the strong law of large numbers [9], we have
limt→∞Mi(t)t=0,i=1,2. |
On the other hand, by (3.10) and (ⅱ), we can derive that limt→∞lnx1(t)t=0. By comparison method (using Lemma 2.1 again), then
limt→∞1t∫t0x2(s)ds=A21¯η1+A11¯η2A12A21+A11A22=Δ2Δ. | (3.12) |
Substituting (3.12) into (3.10) and using Lemma 2.1 again, we can obtain that
limt→∞1t∫t0x1(s)ds=¯η1−A12Δ2ΔA11=Δ1Δ. | (3.13) |
Consequently, we have limt→∞1t∫t0xi(s)ds=ΔiΔ,i=1,2. The proof is completed.
Remark 3.1. Compared with (1.1), model (1.2) is more popular and contains (1.1) as its special case. Another difference between (1.1) and (1.2) is that the Lévy jump is considered in (1.2) while it is not considered in (1.1).
Remark 3.2. Theorem 3.1 implies that the regime switching, time delays and Lévy jumps will bring large influence to the dynamics of (1.2). By changing their values, the species of (1.2) may change from persistence in the mean to extinction, and vice versa, which is analyzed and verified well by numerical simulations in Section 6.
For simplicity, we denote the solution of (1.2) with initial data x(θ)=ϕ(θ),α(θ)=ς by x(t,ϕ,ς). Let x(t,ϕ,ς) be the R2+×S−valued stochastic Markovian process. Let B⊆R2+ be a Borel measurable set and D⊆S, and we denote the transmission probability of the event {x(t,ϕ,ς)∈B×D} by P(t,ϕ,ς,B×D), that is, P(t,ϕ,ς,B×D)=∑u∈D∫BP(t,ϕ,ς,dx×{u}). Denote by P(R2+,S) all the probability measures on R2+×S, and for any two P1,P2∈P(R2+,S), we define the metric dL as follows:
dL(P1,P2)=supf∈L|N∑k=1f(x,k)P1(dx,k)−N∑k=1f(x,k)P2(dx,k)|, | (4.1) |
where L={f:C([−τ,0],R2+)×S→R:|f(x1,k)−f(x2,˜k)≤|x1−x2|+|k−˜k|,|f(⋅,⋅)|≤1}.
Definition 4.1 ([9]). The process x(t,ϕ,ς) is said to be asymptotically stable in distribution if there exists a probability measure u(⋅×⋅) on R2+×S such that the transmission probability P(t,ϕ,ς,dx×{k}) of x(t,ϕ,ς) converges weakly to u(dx×{k}) as t→∞ for every (ϕ,ς)∈C([−τ,0],R2+)×S. System (1.2) is said to be asymptotically stable in distribution if the solution x(t,ϕ,ς) of (1.2) is asymptotically stable in distribution.
Lemma 4.1 ([30]). Let f(t) be a nonnegative function defined on [0,∞) such that f(t) is integrable on [0,∞) and is uniformly continuous on [0,∞), then limt→∞f(t)=0.
For the need of discussion, we give the following technical assumption.
Assumption 3. aii−∫0−τiidμii(θ)−∫0−τjidμji(θ)>aji,i,j=1,2,i≠j.
Assumption 3 means that under the effect of time delays, the intraspecific competition rate is still greater than the interaction rate between different species.
Theorem 4.1. Letx(t)=x(t,ϕ,ς) and ˜x(t)=˜x(t,φ,ς) be two solutions of (1.2) withinitial value x(θ)=ϕ,˜x(θ)=φ and α(θ)=ς, respectively.If Assumption 3 holds, then we have
limt→∞E|xi(t)−˜xi(t)|=0,i=1,2. |
Proof. Define Lyapunov function Vi(t)=|lnxi(t)−ln˜xi(t)|,t≥0,i=1,2. We calculate the right differential of V1(t) along the solutions of (1.2), then
LV1(t)=sgn(x1(t)−˜x1(t))(−a11(x1(t)−˜x1(t))−∫0−τ11(x1(t+θ)−˜x1(t+θ))dμ11(θ)−a12(x2(t)−˜x2(t))−∫0−τ12(x2(t+θ)−˜x2(t+θ))dμ12(θ))≤−a11|x1(t)−˜x1(t)|+a12|x2(t)−˜x2(t)|+∫0−τ11|x1(t+θ)−˜x1(t+θ)|dμ11(θ)+∫0−τ12|x2(t+θ)−˜x2(t+θ)|dμ12(θ). |
Similarly, we have
LV2(t)≤−a22|x2(t)−˜x2(t)|+a21|x1(t)−˜x1(t)|+∫0−τ21|x1(t+θ)−˜x1(t+θ)|dμ21(θ)+∫0−τ22|x2(t+θ)−˜x2(t+θ)|dμ22(θ). |
Define V3(t) as follows:
V3(t)=∫0−τ11∫tt+θ|x1(s)−˜x1(s)|dsdμ11(θ)+∫0−τ12∫tt+θ|x2(s)−˜x2(s)|dsdμ12(θ)+∫0−τ21∫tt+θ|x1(s)−˜x1(s)|dsdμ21(θ)+∫0−τ22∫tt+θ|x2(s)−˜x2(s)|dsdμ22(θ). |
Let V(t)=V1(t)+V2(t)+V3(t), then by Assumption 3, after computation, we have
LV(t)≤−(a11−∫0−τ11dμ11(θ)−∫0−τ21dμ21(θ)−a21)|x1(t)−˜x1(t)|−(a22−∫0−τ21dμ21(θ)−∫0−τ22dμ22(θ)−a12)|x2(t)−˜x2(t)|<0. |
Therefore,
0≤EV(t)≤EV(0)−(a11−∫0−τ11dμ11(θ)−∫0−τ21dμ21(θ)−a21)E|x1(t)−˜x1(t)|−(a22−∫0−τ21dμ21(θ)−∫0−τ22dμ22(θ)−a12)E|x2(t)−˜x2(t)|. |
Hence,
E|xi(t)−˜xi(t)|≤EV(0)aii−∫0−τiidμii(θ)−∫0−τijdμij(θ)∈L1[0,∞) |
for i,j=1,2,i≠j.
On the other hand, by (1.2) we have,
xi(t)=xi(0)+∫t0fi(xi(s),α(s))ds+∫t0gi(xi(s),α(s))dB(s)+∫t0∫Zhi(xi(s),α(s),u)˜N(ds,du), | (4.2) |
where
fi(xi,α(s))=xi(ri(α(s))−hi−aiixi−∫0−τiixi(s+θ)dμii(θ)−aijxj−∫0−τijxj(s+θ)dμij(θ)), |
gi(xi(s),α(s))=xi(s)σi(α(s)),hi(xi(s),α(s),u)=xi(s)γi(α(s),u),i,j=1,2,i≠j. |
Taking expectation from both sides of (4.2), then
Exi(t)=xi(0)+∫t0(ri(α(s))−hi)E(xi)−aiiE(x2i)−∫0−τiiE(xi)E(xi(s+θ))dμii(θ)−aijE(xi)E(xj)−∫0−τijE(xi)E(xj(s+θ))dμij(θ)ds. |
Consequently, E(xi(t)) is continuously differentiable with respect to t. Moreover,
dE(xi(t))dt≤(ri(α(s))−hi)E(xi(t))≤K. |
That is to say, E(xi(t)) is uniformly continuous. An application of Lemma 4.1 gives limt→∞E|xi(t)−˜xi(t)|=0. This completes the proof.
Remark 4.1. Theorem 4.1 will be applied later to prove the stability in distribution of (1.2). Assumption 3 is necessary in our proof. If without S-type time delays and Lévy jumps, Reference [14] implies that Assumption 3 is unnecessary and may be dropped.
Lemma 4.2. For any compact subset B⊆R2+ and (ϕ,ς)∈C([−τ,0],B)×S, the family of transmission probability of the solutionP(t,ϕ,ς,dx×{u}) is tight.
Proof. For (4.2), by use of the Hölder inequality and the moment inequality of stochastic integrals, there exist k=1,2,... such that
E[sup(k−1)δ≤s≤kδ|xi|p]≤4p−1{E|∫kδ(k−1)δfi(xi,α(s))ds|p+Esup(k−1)δ≤s≤kδ|∫kδ(k−1)δgi(xi,α(s))ds|p+|xi((k−1)δ)|p+Esup(k−1)δ≤s≤kδ|∫kδ(k−1)δ∫Zhi(xi,α(s))˜N(ds,du)|p}. | (4.3) |
By Lemma 2.2 (i.e., E|xi(t)|p≤K(p)), then
E|∫kδ(k−1)δfi(xi,α(s))ds|p≤E[δpsup(k−1)δ≤s≤kδxpi(ri(α(s))−hi−aiixi−∫0−τiixi(t+θ)dμii(θ)−aijxj−∫0−τijxj(t+θ)dμij(θ))p]≤2p−1δpE|ri(α(s))−hi−aiixi−∫0−τiixi(t+θ)dμii(θ)−aijxj−∫0−τijxj(t+θ)dμij(θ)|p+2p−1δpE|xi|p≤2p−1δp5p−1{E|ri(α(s))−hi|p+aiiE|xi|p+E|∫0−τiixi(t+θ)dμii(θ)|p+aijE|xj|p+E|∫0−τijxj(t+θ)dμij(θ)|p}+2p−1δpE|xi|p≤10p−1δp[|rui−hi|p+aiiKi(p)+|∫0−τiidμii(θ)|pKi(p)+aijKj(p)+|∫0−τijdμij(θ)|pKj(p)]+2p−1δpKi(p)≜M1(p)δp, | (4.4) |
where M1(p) is a constant. On the other hand, using the Burkholder-Davis-Gundy inequality [19], there exists a constant Cp such that
E[sup(k−1)δ≤s≤kδ|∫kδ(k−1)δg1(x1,α(s))ds|p]≤CpE[∫kδ(k−1)δ|xi(s)σi(α(s))|2ds]p2≤Cpδp2(σu)pE|xi(s)|p≜M2(p)δp2. | (4.5) |
As to the semimartingale part, by applying the Kunita's first inequality [19], there exists a constant D(p) such that
E[sup(k−1)δ≤s≤kδ|∫kδ(k−1)δ∫Zhi(xi(s),α(s),u)˜N(ds,du))|p]≤D(p)E[∫kδ(k−1)δ∫Z|xi(s)γi(α(s),u)|2λ(du)ds]p2+E[∫kδ(k−1)δ∫Z|xi(s)γi(α(s),u)|pλ(du)ds]≤D(p)δp2Kp2K(p)+D(p)δKp2K(p). | (4.6) |
Therefore, for any s∈[0,t], we have
sup(ϕ,ς)∈(C([−τ,0],R2+)×S)E[sup0≤s≤t|x(ϕ,ς,s)|p]<∞. |
That is, the probability measure set P(t,ϕ,ς,dx×{u}) is tight for (ϕ,ς)∈C([−τ,0],R2+)×S. This completes the proof.
Lemma 4.3. Let Assumptions 1–3 hold, then for any (ϕ,ς)∈C([−τ,0],R2+)×S, the transmission probabilityP(t,ϕ,ς,⋅×⋅:t≥0) of the solution of (1.2) is Cauchy in the space P(R2+×S)with the metric dL defined as before.
Proof. Let B={x∈C([−τ,0],R2+):σ≤|x|≤ϱ}, where ϱ is a sufficiently large positive number, and σ is a sufficiently small positive number. By the tightness of the transmission probability of solutions of (1.2), for any ε>0, we have P(t,x,α,BC×S)≤ε,t≥0, where BC=R2+/B.
For any f∈L, t,s>0, and initial value (ϕ,ς)∈C([−τ,0],R2+)×S, by the characteristic of condition expectation, we have
dL(P(t+s,ϕ,ς,⋅×⋅),P(t,ϕ,ς,⋅×⋅))=supf∈L|E[f(x(t+s;ϕ,ς),ας(t+s))]−E[f(x(t;ϕ,ς),ας(t))]|=supf∈L|E[E[f(x(t+s;ϕ,ς),ας(t+s))|FS]]−E[f(x(t;ϕ,ς),ας(t))]|=supf∈L|N∑k=1∫R2+E[f(x(t;φ,k),αk(t))P(s,ϕ,ς,dφ×{k})−E[f(x(t;ϕ,ς),ας(t))]|≤supf∈LN∑k=1∫R2+|E[f(x(t;φ,k),αk(t))−E[f(x(t;ϕ,ς),ας(t))]|P(s,ϕ,ς,dφ×{k})≤supf∈LN∑k=1∫R2+/B|E[f(x(t;φ,k),αk(t))−E[f(x(t;ϕ,ς),ας(t))]|P(s,ϕ,ς,dφ×{k})+supf∈LN∑k=1∫B|E[f(x(t;φ,k),αk(t))−E[f(x(t;ϕ,ς),ας(t))]|P(s,ϕ,ς,dφ×{k}), | (4.7) |
where
supf∈LN∑k=1∫R2+/B|E[f(x(t;φ,k),αk(t))−E[f(x(t;ϕ,ς),ας(t))]|P(s,ϕ,ς,dφ×{k})≤2P(s,ϕ,ς,R2+/B×S)≤2ε. | (4.8) |
By the proof of Theorem 4.6 in [23] (4.22 of page 106), for any ε>0 and sufficiently large t, we have
f(x(t;φ,k),αk(t))−f(x(t;ϕ,ς),ας(t))≤ε. |
Then
supf∈LN∑k=1∫B|E[f(x(t;φ,k),αk(t))−E[f(x(t;ϕ,ς),ας(t))]|P(s,ϕ,ς,dφ×{k})≤supf∈LN∑k=1∫BεP(s,ϕ,ς,dφ×{k})≤ε. | (4.9) |
Hence, dL(P(t+s,ϕ,ς,⋅×⋅),P(t,ϕ,ς,⋅×⋅))≤3ε for s>0 and sufficiently large t, that is, P(t,ϕ,ς,⋅×⋅:t≥0) is Cauchy in the space P(C([−τ,0],R2+)×S). The proof is completed.
Theorem 4.2. Under the conditions of Lemma 4.3, the solution process x(t) of (1.2) is asymptotically stable in distribution.
Proof.By Lemma 4.3, the transmission probability P(t,ϕ,ς,⋅×⋅:t≥0) is Cauchy in the space P(R2+×S) with the metric dL. Hence, for any fixed φ∈C([−τ,0],R2+), P(t,˜x,ς,⋅×⋅:t≥0) is Cauchy in P(R2+×S). Then there exists a probability measure u(⋅×⋅) such that
limt→∞dL(P(t,φ,ς,⋅×⋅),u(⋅×⋅))=0. | (4.10) |
On the other hand, by Theorem 4.1, we have
dLP(t,φ,ς,⋅×⋅),P(t,ϕ,ς,⋅×⋅)))=supf∈L|Ef(x(t,φ,ς)−Ef(x(t,ϕ,ς)|≤supf∈L|E(f(x(t,φ,ς))−f(x(t,ϕ,ς)))|≤E|x(t,φ,ς)−x(t,ϕ,ς)|→0. | (4.11) |
Therefore, by the triangle inequality, we can derive from (4.10) and (4.11) that
limt→∞dL(P(t,ϕ,ς,⋅×⋅),u(⋅×⋅))≤limt→∞dL(P(t,φ,ς,⋅×⋅),u(⋅×⋅))+dL(P(t,φ,ς,⋅×⋅),P(t,ϕ,ς,⋅×⋅))=0. | (4.12) |
Since the weak convergence of probability measure is a metric concept, (4.12) shows that for any initial value (ϕ,ς)∈C([−τ,0],R2+)×S, the probability measure P(t,ϕ,ς,⋅×⋅:t≥0) of the solution of (1.2) converges weakly to the probability measure u(⋅×⋅). By Definition 4.1, then the solution process x(t) of (1.2) is asymptotically stable in distribution. This completes the proof.
Remark 4.2. The asymptotic stability in distribution of species reveals the existence and uniqueness of an invariant probability measure, which is the basis of discussing the optimal harvesting effort in Section 5.
In this section, we consider the optimal harvesting effort (OHE) of (1.2). By [8,27,28,29], the OHE problem is to find a constant h∗=(h∗1,h∗2)T such that both x1 and x2 survive, and Y(h∗)=limt→∞E(h∗ixi(t)) is maximum.
Theorem 5.1. Let λ=(λ1,λ2)T=(A(A−1)T+I)−1ξ, where I is the unit matrix, ξ=(ˉξ1,ˉξ2)T.
(1) If A+(A−1)T is positive semi-definite, and Δ2|hi=λi>0,λi≥0,i=1,2, then the OHE is h∗=λ, and the maximum of the sustainable yield is
Y(h∗)=λTA−1(ξ−λ). |
(2) If the conditions of (1) fail, then there exists no OHE.
Proof. We define Ξ={h=(h1,h2)T∈R2+,Δ2>0}, then for any h∈Ξ, by Theorem 3.1, we have limt→∞1t∫t0xi(t)dt=ΔiΔ,i=1,2. Conversely, if the h∗ of OHE exists, obviously h∗∈Ξ.
(1) By the given conditions, it is clear that λ∈Ξ, that is, the set Ξ is not empty. For any h∈Ξ, by Theorem 3.1, after computation, we have
limt→∞1t∫t0hTx(s)ds=2∑i=1hilimt→∞1t∫t0xi(s)ds=2∑i=1hiΔiΔ=hTA−1(ξ−h). | (5.1) |
Theorem 4.1 implies that the distribution of solutions of (1.2) is asymptotically stable, and hence suppose the stationary probability density is ρ, then
Y(h)=limt→∞E(hixi(t))=limt→∞E(hTx(t))=N∑k=1∫R2+hTxρ(x,k)dx. | (5.2) |
The asymptotical stability in distribution means that there exists a unique invariant measure u. In view of the one-to-one correspondence between the stationary probability density ρ and the invariant measure u (see, e.g.[31], page 105), we have
N∑k=1∫R2+hTxρ(x,k)dx=N∑k=1∫R2+hTxu(dx,k). | (5.3) |
By (5.1)–(5.3), we observe that
Y(h)=hTA−1(ξ−h). | (5.4) |
Clearly, by computing the derivative of the variable h, we have d(Y(h))dh=A−1ξ−(A−1+(A−1)T)h. Denote the solution of d(Y(h))dh=0 by λ, then λ=(AA−1+I)ξ. Further, after computation, we have
dd(Y(h))dhdhT=−(A−1+(A−1)T), |
which is negative semi-definite for all h by the given conditions, then we derive from Theorem 4.1.5 of [32] that λ is a global maximum point of Y(h). That is, if the conditions Δ2|hi=λi>0,λi≥0(i=1,2) hold, then the OHE is h∗=λ, and Y(h∗)=λTA−1(ξ−λ).
(2) Firstly we prove that there is no OHE if Δ2|hi=λi,i=1,2>0, λ1≥0 and λ2≥0, but A−1+(A−1)T is not positive semi-definite. By above conditions, clearly the set Ξ is not empty. Let (ιij)2×2=A−1+(A−1)T, then ι11=2A22Δ>0. It implies that A−1+(A−1)T is not negative semi-definite. Noting that A−1+(A−1)T is not positive semi-definite, hence A−1+(A−1)T is indefinite. The extreme value theory [32] shows that Y(h) has no extreme points. Therefore, there is no OHE.
Secondly, under the conditions Δ2|hi=λi,i=1,2<0, or λ1≤0, or λ2≤0, we prove there is also no OHE. Otherwise, suppose the OHE is ˜h∗=(˜h∗1,˜h∗2)T, then ˜h∗∈Ξ. Thus Δ2|h∗i=˜h∗i>0,˜h∗i≥0,i=1,2. That means ˜h∗i is also the solution of d(Y(h))dh=0, which contradicts with the uniqueness of the solution. This completes the proof.
Remark 5.1. The existence and uniqueness of an invariant probability measure plays a key role to derive (5.4). With the invariant measure, by using the extremum theory, we find the maximum of Y(h), which is very popular and can be applied to resolve some similar problems.
In this section, by numerical analysis, we give some examples and apply the Milstein method [33] to illustrate our theoretical results, and explore the effects of regime switching, time delays and Lévy jumps on the system dynamics.
For simplicity, we assume that the continuous-time discrete state Markovian chain α(t) takes value in the space S={1,2}, then system (1.2) reduces to the following subsystem (6.1) and (6.2):
{dx1(t)=x1(t−)(r1(1)−h1−a11x1(t−)−∫0−τ11x1(t+θ)dμ11(θ)−a12x2(t−)−∫0−τ12x2(t+θ)dμ12(θ))dt+σ1(1)x1(t−)dB1(t)+∫Zγ1(u,1)x1(t−)˜N(dt,du),dx2(t)=x2(t−)(r2(1)−h2+a21x1(t−)+∫0−τ21x1(t+θ)dμ21(θ)−a22x2(t−)−∫0−τ22x2(t+θ)dμ22(θ))dt+σ2(1)x2(t−)dB2(t)+∫Zγ2(u,1)x2(t−)˜N(dt,du), | (6.1) |
and
{dx1(t)=x1(t−)(r1(2)−h1−a11x1(t−)−∫0−τ11x1(t+θ)dμ11(θ)−a12x2(t−)−∫0−τ12x2(t+θ)dμ12(θ))dt+σ1(2)x1(t−)dB1(t)+∫Zγ1(u,2)x1(t−)˜N(dt,du),dx2(t)=x2(t−)(r2(2)−h2+a21x1(t−)+∫0−τ21x1(t+θ)dμ21(θ)−a22x2(t−)−∫0−τ22x2(t+θ)dμ22(θ))dt+σ2(2)x2(t−)dB2(t)+∫Zγ2(u,2)x2(t−)˜N(dt,du). | (6.2) |
For (6.1) and (6.2), unless otherwise stated, we always take a11=0.8,a12=0.6,a21=0.4,a22=0.7, hi=0,σ2i=0.2,∫0−τiidμii(θ)=0.3, ∫0−τijdμij(θ)=0.2,γi(1)=γi(2)=0.4, i=1,2,i≠j,Z=[0,∞), and λ(Z)=1. Then A11=1.1,A12=0.8,−A21=−0.6,A22=1,Δ=1.58.
For (6.1), let r1(1)=0.06,r2(1)=−0.02. It is easy to verify that Assumptions 1–3 hold, and η1(1)=−0.1035<0. Theorem 3.1 implies both x1 and x2 are extinct (see Figure 1 (a)).
For (6.2), let r1(2)=0.7,r2(2)=−0.05. Similarly Assumptions 1–3 hold, and η1(2)=0.5365>0,η2(2)=−0.2135, Δ1=0.7073,Δ2=0.087>0.
By Theorem 3.1 again, we can obtain that both x1 and x2 are persistent in the mean, and
limt→∞1t∫t0x1(s)ds=Δ1Δ=0.4477,limt→∞1t∫t0x2(s)ds=Δ2Δ=0.0551. |
Figure 1 (b) reveals it clearly.
Next we will study the effect of regime switching, time delays and Lévy jumps on the extinction and persistence in the mean of all species, respectively. We discuss in three cases.
● Case (ⅰ) The effect of switching π.
(1) Let the stationary distribution π=(0.1,0.9). By computation, then
ˉη1=0.1∗(−0.1035)+0.9∗0.5365=0.4725,ˉη2=0.1∗(−0.1835)+0.9∗(−0.2135)=−0.2105, |
and
limt→∞1t∫t0x1(s)ds=Δ1Δ=0.4056,limt→∞1t∫t0x2(s)ds=Δ2Δ=0.0328. |
That is, both x1(t) and x2(t) are persistent in the mean. We call it the "persistent case". Figure 2 (a) is the long-term behaviours and Figure 2 (b) is the probability densities of x1(t) and x2(t), respectively. Figure 3 gives the time series and histogram of Markov chain α(t) with stationary distribution π=(0.1,0.9).
For the persistent case, Theorem 5.1 shows that there exists OHE, and
h∗=λ=(A(A−1)T+I)−1¯η=(0.3258,0.0532)T. |
The maximum of the sustainable yield is
Y(h∗)=λTA−1(ξ−λ)=0.0669. |
By numerical simulation, we get Figure 2 (c). In Figure 2 (c), the red line represents the yield Y=0.0669 with λ=(0.3258,0.0532)T, the green line represents the yield Y1=0.0484 with λ1=(0.1558,0.0632)T, the blue line represents the yield Y2=0.0601 with λ2=(0.4258,0.0332)T, respectively. Obviously, the maximum of sustainable yield is Y(h∗)=0.0669. Figure 2 (c) verify it well. For the interpretation of the references to color, readers are referred to the web version of the article.
(2) Let π=(0.9,0.1). Then ˉη1=0.9∗(−0.1035)+0.1∗0.5365=−0.0395<0. Theorem 3.1 implies that both x1(t) and x2(t) are to be extinct. That is, the switching π leads to the extinction of x1(t) and x2(t) (see Figure 4 (a)).
(3) Let π=(0.4,0.6). Then by computing, we have
ˉη1=0.2805>0andΔ2=−0.0534<0. |
By Theorem 3.1 again, then x2 is extinct and x1 is persistent in the mean, and
limt→∞1t∫t0x1(s)ds=ˉη1A11=0.255. |
It shows the switching π leads to the extinction of x2(t) and the different persistence in the mean of x1(t) (see Figure 4 (b)).
● Case (ⅱ) The effect of time-delays.
For the persistent case, if we take ∫0−τiidμii(θ)=0.87,∫0−τijdμij(θ)=0.3, other parameters are same as before, then ˉη1=0.4725>0,Δ2=−0.0208<0, which leads to the extinction of x2(t) and
limt→∞1t∫t0x1(s)ds=0.2829. |
That is, time delays destroy the persistent case, see Figure 4 (c).
● Case (ⅲ) The effect of Lévy jumps.
For the persistent case, if we take γi(1)=γi(2)=1,i=1,2,i≠j, then ¯η1=−0.4189<0. By Theorem 3.1, both x1 and x2 go to be extinct. It implies that too large Lévy jumps leads to the extinction of x1(t) and x2(t), see Figure 4 (d).
In this paper, we study the dynamics of a stochastic predator-prey system with S-type distributed time delays, regime switching and Lévy jumps. Theorem 3.1 gives the sufficient conditions assuring the extinction and persistence in the mean of each species. Theorem 4.2 shows that (1.2) is asymptotically stable in distribution. Theorem 5.1 gives the optimal harvesting effort. Finally, some examples are given and the effects of regime switching, distributed time delays and Lévy jumps are discussed by numerical analysis.
Figure 4 (a) and (b) imply that different regime switching may lead to very different long-term behaviours of species. Figure 2 (c) shows that the OHE is based on the suitable regime switching. Figure 4 (c) and (d) show that too large delays or Lévy jumps will destroy the persistent case of (1.2), respectively. All these show that regime switching, distributed time delays and Lévy jumps play key role in the dynamics of (1.2).
Further, (1.2) is very popular and contains many researched model as its special cases. Firstly, (1.1) is contained in (1.2), and hence, one can obtain the sufficient conditions of the extinction and persistence in the mean for species of (1.1), that is, Theorems 1 and 2 of Reference [14] are contained in our results. Secondly, if α(t)=1,hi=0,γi=0(i=1,2), then we get the model studied by Wang et.al. [18]. Our result coincides with Theorem 2.2 of [18]. Thirdly, if μii(θ) are constant on [−τ,0],aij=0,i≠j, and
μ12(θ)={b12,−τ1<θ≤0,0,−τ12≤θ≤−τ1,μ21(θ)={b21,−τ2<θ≤0,0,−τ21≤θ≤−τ2, |
then we get the discrete time delays model proposed in Reference [5,8]. Similarly we can obtain Theorem 2.2 in [8] and Lemma 3 in [5]. As Liu et. al. stated in [14], the growth of the ith species at time t is often affected by the abundance of the jth species on the interval [t−τi,t], rather than only on the time t−τi. Hence the S-type distributed delays can fit with some real biological systems better. In above sense, we improve and generalize the obtained conclusions of [5,8,14,18].
However, the switching does not appear in all parameters and the control inputs are all constants, if they are dependent on time t, then the persistence in the mean and the optimal harvesting strategy can not be established and is still unknown. On the other hand, for predator-prey system, if the predator is provided with additional food [34], or the fear of prey induced by predator appears [35], what will happen is very interesting. All these will be our research work in the future.
The author is very grateful to the anonymous referees and the editor for their helpful and suggestive comments. This research was supported by the National Natural Science Foundation of China (Nos. 11861027 and 11871475).
The authors declare that they have no competing interests in this paper.
[1] | X. Mao, Stochastic differential equations and applications, England: Horwood Publishing Limited, 2007. |
[2] | N. Ikeda, S. Watanable, Stochastic differential equations and diffusion processes, New York: North-Holland, 1989. |
[3] |
X. Mao, G. Marion, E. Renshaw, Environmental Brownian noise suppresses explosions in populations dynamics, Stoch. Proc. Appl., 97 (2002), 95–110. doi: 10.1016/S0304-4149(01)00126-0. doi: 10.1016/S0304-4149(01)00126-0
![]() |
[4] |
M. Liu, K. Wang, Q. Wu, Survival analysis of stochastic competitive models in a polluted environment and stochastic competitive exclusion principle, Bull. Math. Biol., 73 (2011), 1969–2012. doi: 10.1007/s11538-010-9569-5. doi: 10.1007/s11538-010-9569-5
![]() |
[5] |
M. Liu, H. Qiu, K. Wang, A remark on a stochastic predator-prey system with time delays, Appl. Math. Lett., 26 (2013), 318–323. doi: 10.1016/j.aml.2012.08.015. doi: 10.1016/j.aml.2012.08.015
![]() |
[6] | R. M. May, Stability and complexity in model ecosystems, Princeton: Princeton University Press, 1973. |
[7] | R. Z. Khas'minskii, Stochastic stability of differential equations, Netherlands: Sijthoff Noordhoff, Alphen aan den Rijn, 1980. |
[8] |
M. Liu, Optimal harvesting policy of a stochastic predator-prey model with time delay, Appl. Math. Lett., 48 (2015), 102–108. doi: 10.1016/j.aml.2014.10.007. doi: 10.1016/j.aml.2014.10.007
![]() |
[9] | X. Mao, C. Yuan, Stochastic differential equations with markovian switching, London: Imperial College Press, 2006. |
[10] | Q. Luo, X. Mao, Stochastic population dynamics under regime switching, J. Math. Anal. Appl., 334 (2007), 69–84. |
[11] | W. Kong, Y. Shao, Long-time behaviours of a stochastic predator-prey system with Holling-type Ⅱ functional response and regime switching, J. Math., 2021 (2021), 6672030. |
[12] |
J. Bao, J. Shao, Permanence and extinction of regime-switching predator-prey models, SIAM J. Math. Anal., 48 (2016), 725–739. doi: 10.1137/15M1024512. doi: 10.1137/15M1024512
![]() |
[13] |
H. Chen, P. Shi, C. Lim, Stability analysis for neutral stochastic delay systems with Markovian switching, Syst. Contr. Lett., 110 (2017), 38–48. doi: 10.1016/j.sysconle.2017.10.008. doi: 10.1016/j.sysconle.2017.10.008
![]() |
[14] |
M. Liu, X. He, J. Yu, Dynamics of a stochastic regime-switching predator-prey model with harvesting and distributed delays, Nonlinear Anal.-Hybri., 28 (2018), 87–104. doi: 10.1016/j.nahs.2017.10.004. doi: 10.1016/j.nahs.2017.10.004
![]() |
[15] | Y. Kuang, Delay differential equations: With applications in population dynamics, Boston: Academic Press, 1993. |
[16] |
S. Wang, G. Hu, T. Wei, L. Wang, Stability in distribution of a stochastic predator-prey system with S-type distributed time delays, Physica A, 505 (2018), 919–930. doi: 10.1016/j.physa.2018.03.078. doi: 10.1016/j.physa.2018.03.078
![]() |
[17] |
L. Wang, D. Xu, Global asymptotic stability of bidirectional associative memory neural networks with S-type distributed delays, Int. J. Syst. Sci., 33 (2002), 869–877. doi: 10.1080/00207720210161777. doi: 10.1080/00207720210161777
![]() |
[18] | L. Wang, R. Zhang, Y. Wang, Global exponential stability of reaction-diffusion cellular neural networks with S-type distributed time delays, Nonlinear Anal., 10 (2009), 1101–1113. |
[19] | D. Applebaum, Lévy processes and stochastic Calculus, 2nd ed., Lodon: Cambridge University Press, 2009. |
[20] |
M. Liu, K. Wang, Stochastic Lotka-Volterra systems with Lévy noise, J. Math. Anal. Appl., 410 (2014), 750–763. doi: 10.1016/j.jmaa.2013.07.078. doi: 10.1016/j.jmaa.2013.07.078
![]() |
[21] |
X. Zhang, W. Li, M. Liu, K. Wang, Dynamics of a stochastic Holling Ⅱ one-predator two-prey system with jumps, Physica A, 421 (2015), 571–582. doi: 10.1016/j.physa.2014.11.060. doi: 10.1016/j.physa.2014.11.060
![]() |
[22] |
S. Wang, L. Wang, T. Wei, Permanence and asymptotic behaviors of stochastic predator-prey system with Markovian switching and Lévy noise, Physica A, 495 (2018), 294–311. doi: 10.1016/j.physa.2017.12.088. doi: 10.1016/j.physa.2017.12.088
![]() |
[23] |
Y. Zhao, S. Yuan, Stability in distribution of a stochastic hybrid competitive Lotka-Volterra model with Lévy jumps, Chaos, Solit. Fract., 85 (2016), 98–109. doi: 10.1016/j.chaos.2016.01.015. doi: 10.1016/j.chaos.2016.01.015
![]() |
[24] |
Y. Zhang, Q. Zhang, Dynamic behavior in a delayed stage-structured population model with stochastic fluctuation and harvesting, Nonlinear Dynam., 66 (2011), 231–245. doi: 10.1007/s11071-010-9923-z. doi: 10.1007/s11071-010-9923-z
![]() |
[25] |
J. Yu, M. Liu, Stationary distribution and ergodicity of a stochastic food-chain model with Lévy jumps, Physica A, 482 (2017), 14–28. doi: 10.1016/j.physa.2017.04.067. doi: 10.1016/j.physa.2017.04.067
![]() |
[26] |
G. Hu, K. Wang, Stability in distribution of neutral stochastic functional differential equations with Markovian switching, J. Math. Anal. Appl., 385 (2012), 757–769. doi: 10.1016/j.jmaa.2011.07.002. doi: 10.1016/j.jmaa.2011.07.002
![]() |
[27] |
Y. Shao, Y. Chen, B. X. Dai, Dynamical analysis and optimal harvesting of a stochastic three-species cooperative system with delays and Lévy jumps, Adv. Differ. Equ., 2018 (2018), 423. doi: 10.1186/s13662-018-1874-6. doi: 10.1186/s13662-018-1874-6
![]() |
[28] |
G. D. Liu, X. Z. Meng, Optimal harvesting strategy for a stochastic mutualism system in a polluted environment with regime switching, Physica A, 536 (2019), 120893. doi: 10.1016/j.physa.2019.04.129. doi: 10.1016/j.physa.2019.04.129
![]() |
[29] |
M. Liu, C. Z. Bai, Optimal harvesting of a stochastic mutualism model with regime-switching, Appl. Math. Comput., 373 (2020), 125040. doi: 10.1016/j.amc.2020.125040. doi: 10.1016/j.amc.2020.125040
![]() |
[30] | I. Barbalat, Systems d''equations differentielles d''oscillations non lineaires, Rev. Roum. Math. Pures Appl., 4 (1959), 267–270. |
[31] | D. Prato, J. Zabczyk, Ergodicity for infinite dimensional systems, Cambridge: Cambridge University Press, 1996. doi: 10.1017/CBO9780511662829. |
[32] | M. S. Bazaraa, C. M. Shetty, Nonlinear Programming, Wiley, New York: Academic Press, 1979. |
[33] |
D. J. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. doi: 10.1137/S0036144500378302. doi: 10.1137/S0036144500378302
![]() |
[34] |
D. Amartya, G. P. Samanta, Stochastic prey-predator model with additional food for predator, Physica A, 512 (2018), 121–141. doi: 10.1016/j.physa.2018.08.138. doi: 10.1016/j.physa.2018.08.138
![]() |
[35] |
D. Amartya, G. P. Samanta, Modelling the fear effect on a stochastic prey-predator system with additional food for predator, J. Phys. A-Math. Theor., 51 (2018), 465601. doi: 10.1088/1751-8121/aae4c6. doi: 10.1088/1751-8121/aae4c6
![]() |
1. | Meng Gao, Xiaohui Ai, A stochastic Gilpin-Ayala mutualism model driven by mean-reverting OU process with Lévy jumps, 2024, 21, 1551-0018, 4117, 10.3934/mbe.2024182 | |
2. | Nafeisha Tuerxun, Zhidong Teng, Optimal harvesting strategy of a stochastic $ n $-species marine food chain model driven by Lévy noises, 2023, 31, 2688-1594, 5207, 10.3934/era.2023265 | |
3. | Yuanfu Shao, Jinxing Zhao, Dynamics of a stochastic delayed predator-prey system with regime switching and jumps, 2024, 11, 2522-0144, 10.1007/s40687-024-00480-9 | |
4. | Jinxing Zhao, Yuanfu Shao, Stability of a delayed prey–predator model with fear, stochastic effects and Beddington–DeAngelis functional response, 2025, 2025, 2731-4235, 10.1186/s13662-025-03875-2 |
Conditions | Species y1 | Species y2 |
ˉη1<0 | Extinction | Extinction |
ˉη1>0 and Δ2<0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | Extinction |
Δ2>0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | limt→∞1t∫t0y2(s)ds=Δ2A11A22 |
Cases | Conditions | Species x1 | Species x2 |
ⅰ | ˉη1<0 | Extinction | Extinction |
ⅱ | ˉη1>0 and Δ2<0 | limt→∞1t∫t0x1(s)ds=ˉη1A11 | Extinction |
ⅲ | Δ2>0 | limt→∞1t∫t0x1(s)ds=Δ1Δ | limt→∞1t∫t0x1(s)ds=Δ2Δ |
Conditions | Species y1 | Species y2 |
ˉη1<0 | Extinction | Extinction |
ˉη1>0 and Δ2<0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | Extinction |
Δ2>0 | limt→∞1t∫t0y1(s)ds=ˉη1A11 | limt→∞1t∫t0y2(s)ds=Δ2A11A22 |
Cases | Conditions | Species x1 | Species x2 |
ⅰ | ˉη1<0 | Extinction | Extinction |
ⅱ | ˉη1>0 and Δ2<0 | limt→∞1t∫t0x1(s)ds=ˉη1A11 | Extinction |
ⅲ | Δ2>0 | limt→∞1t∫t0x1(s)ds=Δ1Δ | limt→∞1t∫t0x1(s)ds=Δ2Δ |