
We consider a delayed diffusive predator-prey system with nonlocal competition in prey and schooling behavior in predator. We mainly study the local stability and Hopf bifurcation at the positive equilibrium by using time delay as the parameter. We also analyze the property of Hopf bifurcation by center manifold theorem and normal form method. Through the numerical simulation, we obtain that time delay can affect the stability of the positive equilibrium and induce spatial inhomogeneous periodic oscillations of prey and predator's population densities. In addition, we observe that the increase of space area will not be conducive to the stability of the positive equilibrium (u∗,v∗), and may induce the inhomogeneous periodic oscillations of prey and predator's population densities under some values of the parameters.
Citation: Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang. Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior[J]. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128
[1] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[2] | Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215 |
[3] | Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262 |
[4] | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150 |
[5] | Ruizhi Yang, Dan Jin . Dynamics in a predator-prey model with memory effect in predator and fear effect in prey. Electronic Research Archive, 2022, 30(4): 1322-1339. doi: 10.3934/era.2022069 |
[6] | San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045 |
[7] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[8] | Jiani Jin, Haokun Qi, Bing Liu . Hopf bifurcation induced by fear: A Leslie-Gower reaction-diffusion predator-prey model. Electronic Research Archive, 2024, 32(12): 6503-6534. doi: 10.3934/era.2024304 |
[9] | Mengxin He, Zhong Li . Dynamic behaviors of a Leslie-Gower predator-prey model with Smith growth and constant-yield harvesting. Electronic Research Archive, 2024, 32(11): 6424-6442. doi: 10.3934/era.2024299 |
[10] | Maurıicio F. S. Lima, Jaume Llibre . Hopf bifurcation for a class of predator-prey system with small immigration. Electronic Research Archive, 2024, 32(7): 4604-4613. doi: 10.3934/era.2024209 |
We consider a delayed diffusive predator-prey system with nonlocal competition in prey and schooling behavior in predator. We mainly study the local stability and Hopf bifurcation at the positive equilibrium by using time delay as the parameter. We also analyze the property of Hopf bifurcation by center manifold theorem and normal form method. Through the numerical simulation, we obtain that time delay can affect the stability of the positive equilibrium and induce spatial inhomogeneous periodic oscillations of prey and predator's population densities. In addition, we observe that the increase of space area will not be conducive to the stability of the positive equilibrium (u∗,v∗), and may induce the inhomogeneous periodic oscillations of prey and predator's population densities under some values of the parameters.
Predator-prey relationship exists widely in nature, and many scholars explore this relationship between populations by studying predator-prey model [1,2,3,4,5]. In the real world, the schooling behavior occurs for various reasons among both predator and prey population [6]. By schooling behavior, prey can effectively avoid the capture of predators, and predators can increase the success rate of predation. For example, the wolves [7], African wild dogs and lions [8] are famous examples who have the schooling behavior among predator individuals. To reflect this effect in predator, Cosner et al. [9] proposed the following functional response
η(u,v)=Ce0uv1+thCe0uv, |
where u, v, C, e0, and th represent density of prey, density of predator, capture rate, encounter rate, and handling time, respectively. The functional response η(u,v) monotonically increases with respect to the predator. This reflects that the increase in the number of predators will be conducive to the success rate of predation.
The reaction diffusion equation is widely used in many fields, such as vegetation-water models [10,11], bimolecular models [12,13], population models [14,15,16]. By introducing time and space variables, the reaction-diffusion model can better describe the development law of things. Incorporating the group cooperation in predator and the group defense behavior in prey, J. Yang [17] proposed the following reaction diffusion predator-prey model
{∂u(x,t)∂t=D1Δu+ru(1−uK)−Ce0√uv21+th√uv,∂v(x,t)∂t=D2Δv+v(εCe0√u(x,t−τ)v(x,t−τ)1+th√u(x,t−τ)v(x,t−τ)−d),x∈Ω,t>0∂u(x,t)∂ˉν=∂v(x,t)∂ˉν=0,x∈∂Ω,t>0u(x,θ)=u0(x,θ)≥0,v(x,θ)=v0(x,θ)≥0,x∈ˉΩ,θ∈[−τ,0], | (1.1) |
where u(x,t) and v(x,t) represent prey and predator's densities, respectively. r, K, ε, τ and d represent growth rate, environmental capacity, conversion rate, gestation delay and death rate, respectively. The terms √u and √u(t−τ) represent the herd behavior (or group defense behavior) in prey. They studied saddle-node, Hopf and Bogdanov-Takens types of bifurcations, and discussed the effect of diffusion and time delay on this model through numerical simulations [17].
In the model (1.1), the competition in prey is reflected by the term −uK, which supposes this type competition is spatially local. In fact, the resources is limited in nature, and competition within the population always exist. This competition is usually nonlocal. In [18,19], the authors suggested that the consumption of resources in spatial location is related not only to the local population density, but also to the number of nearby population density. Some scholars have studied the predator-prey models with nonlocal competition [20,21,22]. S. Chen et al studied the existence and uniqueness of positive steady states and Hopf bifurcation in a diffusive predator-prey model with nonlocal effect [20]. J. Gao and S. Guo discussed the steady-state bifurcation and Hopf bifurcation in a diffusive predator-prey model with nonlocal effect and Beddington-DeAngelis Functional Response [21]. S. Djilali studied the pattern formation in a diffusive predator-prey model with herd behavior and nonlocal prey competition, and showed rich dynamic phenomena through numerical simulations [23]. These works suggest that the predator-prey models with nonlocal competition will exhibit different dynamic phenomena compared with the model without nonlocal competition, for example the stably spatially inhomogeneous periodic solutions are more likely to appear.
Based on the model (1.1), we assume there is spatially nonlocal competition in prey. Then, we proposed the following model.
{∂u(x,t)∂t=d1Δu+u(1−∫ΩG(x,y)u(y,t)dy)−α√uv21+√uv,∂v(x,t)∂t=d2Δv+v(β√u(t−τ)v(t−τ)1+√u(t−τ)v(t−τ)−γ),x∈Ω,t>0,∂u(x,t)∂ˉν=∂v(x,t)∂ˉν=0,x∈∂Ω,t>0,u(x,θ)=u0(x,θ)≥0,v(x,θ)=v0(x,θ)≥0,x∈ˉΩ,θ∈[−τ,0]. | (1.2) |
The model (1.2) has been changed by ˜t=r˜t, ˜u=˜uK, ˜v=thCe0√Kv, α=1rt2hCe0K3/2, β=εrth and γ=dr, then drop the tilde. ∫ΩG(x,y)u(y,t)dy represents the nonlocal competition effect in prey. We also choose the Newman boundary condition, which is based on the hypothesis that the region is closed and no prey and predator can leave or enter the boundary.
With the scope of our knowledge, there is no work to study the dynamics of the predator-prey model (1.2) with the nonlocal competition in prey, schooling behavior in predator, reaction diffusion and gestation delay, although it seems more realistic. The aim of this paper is to study the effect of time delay and nonlocal competition on the model (1.2). Whether there exist stable spatially inhomogeneous periodic solutions?
The paper is organized as follows. In Section 2, the stability of coexisting equilibrium and existence of Hopf bifurcation are considered. In Section 3, the property of Hopf bifurcation is studied. In Section 4, some numerical simulations are given. In Section 5, a short conclusion is obtained.
For convenience, we choose Ω=(0,lπ). The kernel function G(x,y)=1lπ, which is based on the assumption that the competition strength among prey individuals in the habitat is the same, that is the competition between any two prey is the same. (0,0) and (1,0) are boundary equilibria of model (1.2). The existence of positive equilibria of model (1.2) has been studied in [17], that is
Lemma 2.1. [17]Assume β>γ, then the model (1.2) has
∙ two distinct coexisting equilibria E1=(u1,v1) and E2=(u2,v2) with 0<u1<35<u2<1 when α<αc(β,γ):=6√15β(β−γ)125γ2;
∙ a unique coexisting equilibrium denoted by E3=(u3,v3) when α=αc(β,γ);
∙ no coexisting equilibrium when α>αc(β,γ).
Make the following hypothesis
(H0)β>γ,α≤αc(β,γ). | (2.1) |
If (H0) holds, then model (1.2) has one or two coexisting equilibria. Hereinafter, for brevity, we just denote E∗(u∗,v∗) as coexisting equilibrium. Linearize model (1.2) at E∗(u∗,v∗)
∂u∂t(u(x,t)u(x,t))=D(Δu(t)Δv(t))+L1(u(x,t)v(x,t))+L2(u(x,t−τ)v(x,t−τ))+L3(ˆu(x,t)ˆv(x,t)), | (2.2) |
where
D=(d100d2),L1=(a1a200),L2=(00b1b2),L3=(−u∗000), |
and a1=1−u∗−v2∗α2√u∗(1+√u∗v∗)2, a2=−(2√u∗v∗+u∗v2∗)α(1+√u∗v∗)2<0, b1=v2∗β2√u∗(1+√u∗v∗)2>0, b2=√u∗v∗β(1+√u∗v∗)2>0, ˆu=1lπ∫lπ0u(y,t)dy. The characteristic equation is
λ2+Anλ+Bn+(Cn−b2λ)e−λτ=0,n∈N0, | (2.3) |
where
A0=u∗−a1,B0=0,C0=−b2(u∗−a1)−a2b1,An=(d1+d2)n2l2−a1,Bn=d1d2n4l4−a1d2n2l2,Cn=−b2d1n2l2+a1b2−a2b1,n∈N. | (2.4) |
When τ=0, the characteristic Eq (2.3) is
λ2+(An−b2)λ+Bn+Cn=0,n∈N0, | (2.5) |
where
{A0−b2=−a1+u∗−b2,B0+C0=−b2(u∗−a1)−a2b1,An−b2=(d1+d2)n2l2−a1−b2,Bn+Cn=d1d2n4l4−(a1d2+b2d1)n2l2+a1b2−a2b1,n∈N. | (2.6) |
Make the following hypothesis
(H1)An−b2>0,Bn+Cn>0,forn∈N0. | (2.7) |
Theorem 2.2. For model (1.2), assume τ=0 and (H0) holds. Then E∗(u∗,v∗) is locally asymptotically stable under (H1).
Proof. If (H1) holds, we can obtain that the characteristic root of (2.5) all have negative real parts. Then E∗(u∗,v∗) is locally asymptotically stable.
Let iω (ω>0) be a solution of Eq (2.3), then
−ω2+iωAn+Bn+(Cn−b2iω)(cosωτ−isinωτ)=0. |
We can obtain cosωτ=ω2(b2An+Cn)−BnCnC2n+b22ω2, sinωτ=ω(AnCn+Bnb2−b2ω2)C2n+b22ω2. It leads to
ω4+ω2(A2n−2Bn−b22)+B2n−C2n=0. | (2.8) |
Let z=ω2, then (2.8) becomes
z2+z(A2n−2Bn−b22)+B2n−C2n=0, | (2.9) |
and the roots of (2.9) are z±=12[−Pn±√P2n−4QnRn], where Pn=A2n−2Bn−b22, Qn=Bn+Cn, and Rn=Bn−Cn. If (H0) and (H1) hold, Qn>0(n∈N0). By direct calculation, we have
P0=(a1−u∗)2−b22>0,Pk=(a1−d1k2l2)2+d22n4l4−b22,R0=a2b1+b2(u∗−a1)Rk=d1d2k4l4+(b2d1−a1d2)k2l2+a2b1−a1b2,fork∈N. | (2.10) |
Define
W1={n|Rn<0,n∈N0},W2={n|Rn>0,Pn<0,P2n−4QnRn>0,n∈N},W3={n|Rn>0,P2n−4QnRn<0,n∈N0}, | (2.11) |
and
ω±n=√z±n,τj,±n={1ω±narccos(V(n,±)cos)+2jπ,V(n,±)sin≥0,1ω±n[2π−arccos(V(n,±)cos)]+2jπ,V(n,±)sin<0.V(n,±)cos=(ω±n)2(b2An+Cn)−BnCnC2n+b22(ω±n)2,V(n,±)sin=ω±n(AnCn+Bnb2−b2(ω±n)2)C2n+b22(ω±n)2. | (2.12) |
We have the following lemma.
Lemma 2.3. Assume (H0) and (H1) hold, the following results hold.
∙ Eq (2.3) has a pair of purely imaginary roots±iω+n at τj,+n for j∈N0 and n∈W1.
∙ Eq (2.3) has two pairs of purely imaginary roots±iω±n at τj,±n for j∈N0 and n∈W2.
∙ Eq (2.3) has no purely imaginary root for n∈W3.
Lemma 2.4. Assume (H0) and (H1) hold. Then Re(dλdτ)|τ=τj,+n>0, Re(dλdτ)|τ=τj,−n<0 for n∈W1∪W2 and j∈N0.
Proof. By Eq (2.3), we have
(dλdτ)−1=2λ+An−b2e−λτ(Cn−b2λ)λe−λτ−τλ. |
Then
[Re(dλdτ)−1]τ=τj,±n=Re[2λ+An−b2e−λτ(Cn−b2λ)λe−λτ−τλ]τ=τj,±n=[1C2n+b22ω2(2ω2+A2n−2Bn−b22)]τ=τj,±n=±[1C2n+b22ω2√(A2n−2Bn−b22)2−4(B2n−C2n)]τ=τj,±n. |
Therefore Re(dλdτ)|τ=τj,+n>0, Re(dλdτ)|τ=τj,−n<0.
Denote τ∗=min{τ0n|n∈W1∪W2}. We have the following theorem.
Theorem 2.5. Assume (H0) and (H1) hold, then the following statements are true for model (1.2).
∙ E∗(u∗,v∗) is locally asymptotically stable for τ>0 when W1∪W2=∅.
∙ E∗(u∗,v∗) is locally asymptotically stable for τ∈[0,τ∗) when W1∪W2≠∅.
∙ E∗(u∗,v∗) is unstable for τ∈(τ∗,τ∗+ε) for some ε>0 when W1∪W2≠∅.
∙ Hopf bifurcation occurs at(u∗,v∗) when τ=τj,+n (τ=τj,−n), j∈N0, n∈W1∪W2. The bifurcating periodic solutions are spatially homogeneous when τ=τj,+0 (τ=τj,−0), and spatially inhomogeneous when τ=τj,+n (τ=τj,−n) for n∈N.
By the works [24,25], we study the property of Hopf bifurcation. For fixed j∈N0 and n∈W1∪W2, we denote ˜τ=τj,±n. Let ˉu(x,t)=u(x,τt)−u∗ and ˉv(x,t)=v(x,τt)−v∗. Drop the bar, (1.2) can be written as
{∂u∂t=τ[d1Δu+(u+u∗)(1−1lπ∫lπ0(u(y,t)+u∗)dy)−α√u+u∗(v+v∗)21+√u+u∗(v+v∗)],∂v∂t=τ[d2Δv+(β√u(t−1)+u∗(v(t−1)+v∗)1+√u(t−1)+u∗(v(t−1)+v∗)−γ)(v+v∗)]. | (3.1) |
Rewrite the model (3.1) as
{∂u∂t=τ[d1Δu+a1u+a2v−u∗ˆu+α1u2−uˆu+α2uv+α3v2+α4u3+α5u2v+α6uv2+α7v3]+h.o.t.,∂v∂t=τ[d2Δv+b1u(t−1)+b2v(t−1)+β1u2(t−1)+β2u(t−1)v(t−1)+β3u2(t−1)+β4u3(t−1)+β5u2(t−1)v(t−1)]+β6u(t−1)v2(t−1)+β7v3(t−1)]+h.o.t., | (3.2) |
where α1=v2∗(1+3√u∗v∗)α8u3/2∗(1+√u∗v∗)3, α2=−v∗α√u∗(1+√u∗v∗)3, α3=−√u∗α(1+√u∗v∗)3, α4=−v2∗(1+4√u∗v∗+5u∗v2∗)α16u5/2∗(1+√u∗v∗)4, α5=v∗(1+4√u∗v∗)αu3/2∗(1+√u∗v∗)4, α6=(−1+2√u∗v∗)α2√u∗(1+√u∗v∗)4, α7=u∗α(1+√u∗v∗)4, β1=−v2∗(1+3√u∗v∗)β2u3/2∗(1+√u∗v∗)3, β2=−v∗(−1+√u∗v∗)β2√u∗(1+√u∗v∗)3, β3=−u∗v∗β(1+√u∗v∗)3, β4=v2∗(1+4√u∗v∗+5u∗v2∗)β16u5/2∗(1+√u∗v∗)4, β5=v∗(−1−4√u∗v∗+3u∗v2∗)β2u3/2∗(1+√u∗v∗)4, β6=v∗(−2+√u∗v∗)β2(1+√u∗v∗)4, β7=u3/2∗v∗β(1+√u∗v∗)4.
Define the real-valued Sobolev space X:={(u,v)T:u,v∈H2(0,lπ),(ux,vx)|x=0,lπ=0}, the complexification of X XC:=X⊕iX={x1+ix2|x1,x2∈X}. and the inner product <˜u,˜v>:=∫lπ0¯u1v1dx+∫lπ0¯u2v2dx for ˜u=(u1,u2)T, ˜v=(v1,v2)T, ˜u,˜v∈XC. The phase space C:=C([−1,0],X) is with the sup norm, then we can write ϕt∈C, ϕt(θ)=ϕ(t+θ) or −1≤θ≤0. Denote β(1)n(x)=(γn(x),0)T, β(2)n(x)=(0,γn(x))T, and βn={β(1)n(x),β(2)n(x)}, where {β(i)n(x)} is an an orthonormal basis of X. We define the subspace of C as Bn:=span{<ϕ(⋅),β(j)n>β(j)n|ϕ∈C,j=1,2}, n∈N0. There exists a 2×2 matrix function ηn(σ,˜τ) −1≤σ≤0, such that −˜τDn2l2ϕ(0)+˜τL(ϕ)=∫0−1dηn(σ,τ)ϕ(σ) for ϕ∈C. The bilinear form on C∗×C is defined by
(ψ,ϕ)=ψ(0)ϕ(0)−∫0−1∫σξ=0ψ(ξ−σ)dηn(σ,˜τ)ϕ(ξ)dξ, | (3.3) |
for ϕ∈C, ψ∈C∗. Define τ=˜τ+μ, then the system undergoes a Hopf bifurcation at (0,0) when μ=0, with a pair of purely imaginary roots ±iωn0. Let A denote the infinitesimal generators of semigroup, and A∗ be the formal adjoint of A under the bilinear form (3.3). Define the following function
δ(n0)={1n0=0,0n0∈N. | (3.4) |
Choose ηn0(0,˜τ)=˜τ[(−n20/l2)D+L1+L3δ(nn0)], ηn0(−1,˜τ)=−˜τL2, ηn0(σ,˜τ)=0 for −1<σ<0. Let p(θ)=p(0)eiωn0˜τθ(θ∈[−1,0]), q(ϑ)=q(0)e−iωn0˜τϑ(ϑ∈[0,1]) be the eigenfunctions of A(˜τ) and A∗ corresponds to iωn0˜τ respectively. We can choose p(0)=(1,p1)T, q(0)=M(1,q2), where p1=1a2(iωn0+d1n20/l2−a1+u∗δ(n0)), q2=a2/(iωn0−b2eiτωn0+d2n2l2), and M=(1+p1q2+˜τq2(b1+b2p1)e−iωn0˜τ)−1. Then (3.1) can be rewritten in an abstract form
dU(t)dt=(˜τ+μ)DΔU(t)+(˜τ+μ)[L1(Ut)+L2U(t−1)+L3ˆU(t)]+F(Ut,ˆUt,μ), | (3.5) |
where
F(ϕ,μ)=(˜τ+μ)(α1ϕ1(0)2−ϕ1(0)ˆϕ1(0)+α2ϕ1(0)ϕ2(0)+α3ϕ2(0)2+α4ϕ31(0)+α5ϕ21(0)ϕ2(0)+α6ϕ1(0)ϕ22(0)+α7ϕ32(0)β1ϕ21(−1)+β2ϕ1(−1)ϕ2(−1)+β3ϕ22(−1)+β4ϕ31(−1)+β4ϕ21(−1)ϕ2(−1)+β6ϕ1(−1)ϕ22(−1)+β7ϕ32(−1)) | (3.6) |
respectively, for ϕ=(ϕ1,ϕ2)T∈C and ˆϕ1=1lπ∫lπ0ϕdx. Then the space C can be decomposed as C=P⊕Q, where P={zpγn0(x)+ˉzˉpγn0(x)|z∈C}, Q={ϕ∈C|(qγn0(x),ϕ)=0and(ˉqγn0(x),ϕ)=0}. Then, model (3.6) can be rewritten as Ut=z(t)p(⋅)γn0(x)+ˉz(t)ˉp(⋅)γn0(x)+ω(t,⋅) and ^Ut=1lπ∫lπ0Utdx, where
z(t)=(qγn0(x),Ut),ω(t,θ)=Ut(θ)−2Re{z(t)p(θ)γn0(x)}. | (3.7) |
then, we have ˙z(t)=iω)n0˜τz(t)+ˉq(0)<F(0,Ut),βn0>. There exists a center manifold C0 and ω can be written as follow near (0,0).
ω(t,θ)=ω(z(t),ˉz(t),θ)=ω20(θ)z22+ω11(θ)zˉz+ω02(θ)ˉz22+⋯. | (3.8) |
Then, restrict the system to the center manifold is ˙z(t)=iωn0˜τz(t)+g(z,ˉz). Denote g(z,ˉz)=g20z22+g11zˉz+g02ˉz22+g21z2ˉz2+⋯. By direct computation, we have
g20=2˜τM(ς1+q2ς2)I3,g11=˜τM(ϱ1+q2ϱ2)I3,g02=ˉg20, |
g21=2˜τM[(κ11+q2κ21)I2+(κ12+q2κ22)I4], |
where I2=∫lπ0γ2n0(x)dx, I3=∫lπ0γ3n0(x)dx, I4=∫lπ0γ4n0(x)dx, ς1=−δn+α1+α2ξ+α3ξ2, ς2=e−2iτωn(β1+ξ(β2+β3ξ)), ϱ1=14(2α1−2δn+α2ˉξ+α2ξ+2α3ˉξξ), ϱ2=14(2β1+2β3ˉξξ+β2(ˉξ+ξ)), κ11=2W(1)11(0)(−1+2α1−δn+α2ξ)+2W(2)11(0)(α2+2α3ξ)+W(1)20(0)(−1+2α1−δn+α2ˉξ)+W(2)20(0)(α2+2α3ˉξ), κ12=12(3α4+α5(ˉξ+2ξ)+ξ(2α6ˉξ+α6ξ+3α7ˉξξ)), κ21=2e−iτωnW(1)11(−1)(2β1+β2ξ)+2e−iτωnW(2)11(−1)(β2+2β3ξ)+eiτωnW(1)20(−1)(2β1+β2ˉξ)+eiτωnW(2)20(−1)(β2+2β3ˉξ), κ22=12e−iτωn(3β4+β5(ˉξ+2ξ)+ξ(2β6ˉξ+β6ξ+3β7ˉξξ)).
Now, we compute W20(θ) and W11(θ) for θ∈[−1,0] to give g21. By (3.7), we have
˙ω=˙Ut−˙zpγn0(x)−˙ˉzˉpγn0(x)=Aω+H(z,ˉz,θ), | (3.9) |
where
H(z,¯z,θ)=H20(θ)z22+H11(θ)z¯z+H02(θ)¯z22+⋯. | (3.10) |
Compare the coefficients of (3.8) with (3.9), we have
(A−2iωn0˜τI)ω20=−H20(θ),Aω11(θ)=−H11(θ). | (3.11) |
Then, we have
ω20(θ)=−g20iωn0˜τp(0)eiωn0˜τθ−ˉg023iωn0˜τˉp(0)e−iωn0˜τθ+E1e2iωn0˜τθ,ω11(θ)=g11iωn0˜τp(0)eiωn0˜τθ−ˉg11iωn0˜τˉp(0)e−iωn0˜τθ+E2, | (3.12) |
where E1=∑∞n=0E(n)1, E2=∑∞n=0E(n)2,
E(n)1=(2iωn0˜τI−∫0−1e2iωn0˜τθdηn0(θ,ˉτ))−1<˜F20,βn>,E(n)2=−(∫0−1dηn0(θ,ˉτ))−1<˜F11,βn>,n∈N0, | (3.13) |
<˜F20,βn>={1lπˆF20,n0≠0,n=0,12lπˆF20,n0≠0,n=2n0,1lπˆF20,n0=0,n=0,0,other,<˜F11,βn>={1lπˆF11,n0≠0,n=0,12lπˆF11,n0≠0,n=2n0,1lπˆF11,n0=0,n=0,0,other, |
and ˆF20=2(ς1,ς2)T, ˆF11=2(ϱ1,ϱ2)T.
Thus, we can obtain
c1(0)=i2ωn˜τ(g20g11−2|g11|2−|g02|23)+12g21,μ2=−Re(c1(0))Re(λ′(˜τ)),T2=−1ωn0˜τ[Im(c1(0))+μ2Im(λ′(τjn))],β2=2Re(c1(0)). | (3.14) |
Theorem 3.1. For any critical value τjn (n∈S,j∈N0), we have the following results.
∙ When μ2>0 (resp.<0), the Hopf bifurcation is forward (resp. backward).
∙ When β2<0 (resp. >0), the bifurcating periodic solutions on the center manifold are orbitally asymptotically stable(resp. unstable).
∙ When T2>0 (resp. T2<0), the period increases (resp. decreases).
To verify our theoretical results, we give the following numerical simulations. Fix parameters
α=1.07,γ=0.2,l=2,d1=1,d2=1. | (4.1) |
The bifurcation diagram of model (1.2) with parameter β is given in Figure 1. We can see that with the increase of parameter β, the stable region of positive equilibrium (u∗,v∗) will decrease.
Especially, fix β=0.595, we can obtain E1≈(0.5362,0.6914) and E2≈(0.6616,0.6225) are two positive equilibria. It is easy to obtain that E2 is always unstable. Then we mainly consider the stability of E1. It can be verified that (H1) holds. By direct calculation, we have τ∗=τ0,+1≈0.6271<τ0,+0≈5.7949. When τ=τ∗, we have μ2≈147.6936, β2≈−6.770 and T2≈48.7187, then E1 is locally asymptotically stable for τ<τ∗ (shown in Figure 2). And the stable inhomogeneous periodic solutions exists for τ>τ∗ (shown in Figure 3). To compare our result with the work in [17], we give the numerical simulations of model (1.2) without nonlocal competition same with the model in [17] under the same parameter τ=4 in Figure 4. We can see that nonlocal competition is the key to the existence of stable inhomogeneous periodic solutions.
To consider the effect of space length on the stability of the positive equilibrium (u∗,v∗), we give the bifurcation diagram of model (1.2) with parameter l (Figure 5) as other parameters fixed in (4.1) and β=0.595. We can see that when the parameter l smaller than the critical value, stable region of the positive equilibrium (u∗,v∗) remains unchanged. This means that the spatial diffusion will not affect the stability of the positive equilibrium (u∗,v∗). When the parameter l is larger than the critical value, increasing of parameter l will cause the stable region of positive equilibrium (u∗,v∗) decrease. This means that the increase of space area will not be conducive to the stability of the positive equilibrium (u∗,v∗), and the inhomogeneous periodic oscillations of prey and predator's population densities may occur.
In this paper, we study a delayed diffusive predator-prey system with nonlocal competition and schooling behavior in prey. By using time delay as parameter, we study the local stability of the positive equilibrium and Hopf bifurcation at the positive equilibrium. We also analyze the property of Hopf bifurcation by center manifold theorem and normal form method. Through numerical simulation, we consider the effect of nonlocal competition on the model (1.2). Our results suggest that time delay can affect the stability of the positive equilibrium. When time delay is smaller than the critical value, the positive equilibrium is locally stable, and becomes unstable when time delay larger than the critical value. Then the prey and predator's population densities will oscillate periodically. But under the same parameters, spatial inhomogeneous periodic oscillations of prey and predator's population densities will appear in the model with nonlocal competition, and prey and predator's population densities will tend to the positive equilibrium in the model without nonlocal competition. This means that time delay can induce spatial inhomogeneous periodic oscillations in the predator-prey model with the nonlocal competition term, which is different from the model without the nonlocal competition term. In addition, we obtain that the increase of space area will not be conducive to the stability of the positive equilibrium (u∗,v∗), and may induce the inhomogeneous periodic oscillations of prey and predator's population densities under some parameters.
This research is supported by the Fundamental Research Funds for the Central Universities (Grant No. 2572022BC01), Postdoctoral program of Heilongjiang Province (No. LBH-Q21060), and College Students Innovations Special Project funded by Northeast Forestry University (No. 202210225160).
The authors declare there is no conflicts of interest.
[1] |
X. Zhang, Q. An, L. Wang, Spatiotemporal dynamics of a delayed diffusive ratio-dependent predator-prey model with fear effect, Nonlinear Dyn., 105 (2021), 3775–3790. https://doi.org/10.1007/s11071-021-06780-x doi: 10.1007/s11071-021-06780-x
![]() |
[2] |
R. Yang, C. Zhang, Dynamics in a diffusive predator–prey system with a constant prey refuge and delay, Nonlinear Anal. Real World Appl., 31 (2016), 1–22. https://doi.org/10.1016/j.nonrwa.2016.01.005 doi: 10.1016/j.nonrwa.2016.01.005
![]() |
[3] |
S. Djilali, S. Bentout, Pattern formations of a delayed diffusive predator-prey model with predator harvesting and prey social behavior, Math. Methods Appl. Sci., 30 (2021), 9128–9142. https://doi.org/10.1002/mma.7340 doi: 10.1002/mma.7340
![]() |
[4] |
R. Yang, X. Zhao, Y. An, Dynamical analysis of a delayed diffusive predator-prey model with additional food provided and anti-predator behavior, Mathematics, 10 (2022), 469. https://doi.org/10.3390/math10030469 doi: 10.3390/math10030469
![]() |
[5] |
R. Yang, D. Jin, W. Wang, A diffusive predator-prey model with generalist predator and time delay, AIMS Mathematics, 7 (2022), 4574–4591. https://doi.org/10.3934/math.2022255 doi: 10.3934/math.2022255
![]() |
[6] |
D. I. Rubenstein, On predation, competition, and the advantages of group living, Social Behav., 3 (1978), 469. https://doi.org/10.1007/978-1-4684-2901-5_9 doi: 10.1007/978-1-4684-2901-5_9
![]() |
[7] |
P. A. Schmidt, D. W. Macdonald, Wolf pack size and food acquisition, Am. Nat., 150 (1997), 513–517. https://doi.org/10.1086/286079 doi: 10.1086/286079
![]() |
[8] |
D. Scheel, C. Packer, Group hunting behaviour of lions: A search for cooperation, Anim. Behav., 41 (1991), 697–709. https://doi.org/10.1016/S0003-3472(05)80907-8 doi: 10.1016/S0003-3472(05)80907-8
![]() |
[9] |
C. Cosner, D. L. Deangelis, J. S. Ault, D. B. Olson, Effects of spatial grouping on the functional response of predators, Theor. Popul. Biol., 56 (1999), 65–75. https://doi.org/10.1006/tpbi.1999.1414 doi: 10.1006/tpbi.1999.1414
![]() |
[10] | J. Li, G. Sun, Z. Guo, Bifurcation analysis of an extended Klausmeier-Gray-Scott model with infiltration delay, Stud. Appl. Math., 2022. https://doi.org/10.1111/sapm.12482 |
[11] |
Q. Xue, C. Liu, L. Li, G. Sun, Z. Wang, Interactions of diffusion and nonlocal delay give rise to vegetation patterns in semi-arid environments, Appl. Math. Comput., 399 (2021), 126038. https://doi.org/10.1016/j.amc.2021.126038 doi: 10.1016/j.amc.2021.126038
![]() |
[12] |
X. P. Yan, C. H. Zhang, Turing instability and formation of temporal patterns in a diffusive bimolecular model with saturation law, Nonlinear Anal. Real World Appl., 43 (2018), 54–77. https://doi.org/10.1016/j.nonrwa.2018.02.004 doi: 10.1016/j.nonrwa.2018.02.004
![]() |
[13] |
F. Yi, J. Liu, J. Wei, Spatiotemporal pattern formation and multiple bifurcations in a diffusive bimolecular model, Nonlinear Anal. Real World Appl., 11 (2010), 3770–3781. https://doi.org/10.1016/j.nonrwa.2010.02.007 doi: 10.1016/j.nonrwa.2010.02.007
![]() |
[14] |
R. Yang, D. Jin, Dynamics in a predator-prey model with memory effect in predator and fear effect in prey, Electron. Res. Arch., 30 (2022), 1322–1339. https://doi.org/10.3934/era.2022069 doi: 10.3934/era.2022069
![]() |
[15] |
R. Yang, L. Wang, D. Jin, Hopf bifurcation analysis of a diffusive nutrient-phytoplankton model with time delay, Axioms, 11 (2022), 56. https://doi.org/10.3390/axioms11020056 doi: 10.3390/axioms11020056
![]() |
[16] |
R. Yang, Q. Song, Y. An, Spatiotemporal dynamics in a predator–prey mModel with functional response increasing in both predator and prey densities, Mathematics, 10 (2022), 17. https://doi.org/10.3390/math10010017 doi: 10.3390/math10010017
![]() |
[17] |
J. Yang, S. Yuan, T. Zhang, Complex dynamics of a predator-prey system with herd and schooling behavior: With or without delay and diffusion, Nonlinear Dyn., 104 (2021), 1709–1735. https://doi.org/10.1007/s11071-021-06343-0 doi: 10.1007/s11071-021-06343-0
![]() |
[18] |
N. F. Britton, Aggregation and the competitive exclusion principle, J. Theor. Biol., 136 (1989), 57–66. https://doi.org/10.1016/S0022-5193(89)80189-4 doi: 10.1016/S0022-5193(89)80189-4
![]() |
[19] |
J. Furter, M. Grinfeld, Local vs. non-local interactions in population dynamics, J. Math. Biol., 27 (1989), 65–80. https://doi.org/10.1007/BF00276081 doi: 10.1007/BF00276081
![]() |
[20] |
S. Wu, Y. Song, Spatiotemporal dynamics of a diffusive predator-prey model with nonlocal effect and delay, Commun. Nonlinear Sci. Numer. Simul., 89 (2020), 105310. https://doi.org/10.1016/j.cnsns.2020.105310 doi: 10.1016/j.cnsns.2020.105310
![]() |
[21] |
J. Gao, S. Guo, Patterns in a modified Leslie-Gower model with Beddington-DeAngelis functional response and nonlocal prey competition, Int. J. Bifurcation Chaos, 30 (2020), 2050074. https://doi.org/10.1142/S0218127420500741 doi: 10.1142/S0218127420500741
![]() |
[22] |
Y. Liu, D. Duan, B. Niu, Spatiotemporal dynamics in a diffusive predator-prey model with group defense and nonlocal competition, Appl. Math. Lett., 103 (2020), 106175. https://doi.org/10.1016/j.aml.2019.106175 doi: 10.1016/j.aml.2019.106175
![]() |
[23] |
S. Djilali, Pattern formation of a diffusive predator-prey model with herd behavior and nonlocal prey competition, Math. Methods Appl. Sci., 43 (2020), 2233–2250. https://doi.org/10.1002/mma.6036 doi: 10.1002/mma.6036
![]() |
[24] | J. Wu, Theory and Applications of Partial Functional Differential Equations, Springer New York, NY, 1996. https://doi.org/10.1007/978-1-4612-4050-1 |
[25] | B. D. Hassard, N. D. Kazarinoff, Y. H. Wan, Theory and Applications of Hopf Bifurcation, Cambridge University Press, Cambridge, 1981. http://dx.doi.org/10.1090/conm/445 |
1. | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang, The dynamics of a delayed predator-prey model with square root functional response and stage structure, 2024, 32, 2688-1594, 3275, 10.3934/era.2024150 |