Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Pattern formation in a ratio-dependent predator-prey model with cross diffusion

  • Received: 25 August 2022 Revised: 21 November 2022 Accepted: 05 December 2022 Published: 15 December 2022
  • This paper is focused on a ratio-dependent predator-prey model with cross-diffusion of quasilinear fractional type. By applying the theory of local bifurcation, it can be proved that there exists a positive non-constant steady state emanating from its semi-trivial solution of this problem. Further based on the spectral analysis, such bifurcating steady state is shown to be asymptotically stable when the cross diffusion rate is near some critical value. Finally, numerical simulations and ecological interpretations of our results are presented in the discussion section.

    Citation: Qing Li, Junfeng He. Pattern formation in a ratio-dependent predator-prey model with cross diffusion[J]. Electronic Research Archive, 2023, 31(2): 1106-1118. doi: 10.3934/era.2023055

    Related Papers:

    [1] 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
    [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] 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
    [4] 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
    [5] Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128
    [6] 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
    [7] 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
    [8] Yuan Tian, Yang Liu, Kaibiao Sun . Complex dynamics of a predator-prey fishery model: The impact of the Allee effect and bilateral intervention. Electronic Research Archive, 2024, 32(11): 6379-6404. doi: 10.3934/era.2024297
    [9] Ailing Xiang, Liangchen Wang . Boundedness of a predator-prey model with density-dependent motilities and stage structure for the predator. Electronic Research Archive, 2022, 30(5): 1954-1972. doi: 10.3934/era.2022099
    [10] 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
  • This paper is focused on a ratio-dependent predator-prey model with cross-diffusion of quasilinear fractional type. By applying the theory of local bifurcation, it can be proved that there exists a positive non-constant steady state emanating from its semi-trivial solution of this problem. Further based on the spectral analysis, such bifurcating steady state is shown to be asymptotically stable when the cross diffusion rate is near some critical value. Finally, numerical simulations and ecological interpretations of our results are presented in the discussion section.



    This paper investigates the following ratio-dependent predator-prey model with cross-diffusion

    {ut=Δu+u(kubvmu+v),xΩ,t>0,vt=Δ[(1+β1+ρu)v]+v(lv+cumu+v),xΩ,t>0,u=v=0,xΩ,t>0,u(x,0)=u0(x)0,v(x,0)=v0(x)0,x Ω, (1.1)

    where the habitat Ω is a bounded domain in Rn with smooth boundary Ω. The functions u(x,t) and v(x,t) denote the densities of prey and predator at space location x and time t, respectively. The coefficients b,c,k,l and m are all positive constants; β and ρ are non-negative constants. The cross diffusion term Δ(βv1+ρu) indicates that the population pressure of predator species diminishes where the density of prey species is high. The prey and predator interact according to the ratio-dependent functional response which is given by the term uvmu+v when the predators hunt seriously.

    As is known, kinds of biological populations correspond to various peculiar and interesting ecological phenomena. Among them, most of such phenomena appearing in ecology can be described by the reaction-diffusion models. These models typically consist of the diffusion term and the reaction term. Along with the innovative work [1] by Shigesada et al., many mathematicians and ecologists have been dedicated to investigate the population models with cross-diffusion from different mathematical perspectives. Especially for the bounded domain, the global existence in time, the existence and stability of positive steady states have been extensively studied in recent two decades (see [2,3,4,5,6,7,8,9,10] and the references therein). For the prey-predator relationship as Eq (1.1), the cross-diffusion term of this fractional type is consistent with the typical pattern of observed interaction between prey and predator. Biomathematical models with such cross-diffusion terms have attracted tremendous attention of scholars (see [11,12,13,14] and the references therein). Considering the saturation of predators for preys, the Holling-Ⅱ functional response uvmu+1 is widely used to describe the interactions of preys and predators. In [12], Wang and Li proved the global bifurcation branch of the positive solution for the prey-predator system with Holling-Ⅱ functional response and cross diffusion. They also investigated the limiting behavior of the steady states when the cross-diffusion approaches infinity. The model was further studied in [13], where the qualitative properties of positive solutions are obtained by applying the Leray-Schauder degree theory and bifurcation argument. Under Neumann boundary condition, Cao et al. [14] have considered the Turing instability in the prey-predator system with cross diffusion when the diffusion coefficients of prey are negative. However, in some situations, especially when predators have to search, share or compete for food, a more suitable prey-predator interaction should be the ratio-dependent functional response uvmu+v. For a more detailed biological description of the ratio-dependent functional response, one can refer to [15,16,17,18,19] and the references therein. As far as we know, there are few studies concerned on the model with both cross-diffusion and ratio-dependent functional response. Recently in virtue of a priori estimates and bifurcation theory, Kumari and Mohan [20] proved the existence and the global bifurcation set of positive steady states to the Eq (1.1) with the bifurcation parameter l.

    In the current work, we focus on the existence and stability of the nonconstant positive steady state for the ratio-dependent prey-predator model (1.1) with cross-diffusion, which are closely related to the pattern formation between predator and prey in ecology. In Section 2, it can be proved that there exists the nonconstant steady states of the nonlinear predator-prey model (1.1) bifurcating from the semi-trivial solution with the bifurcation parameter ρ. In Section 3, we study the stability of such bifurcating steady states by virtue of the spectral analysis. This paper ends with a discussion section containing both numerical simulations and ecological interpretations of our results.

    At the beginning of this section, we introduce some symbols and known results. For p>n, define the Banach spaces

    X=[W2,p(Ω)W1,p0(Ω)]×[W2,p(Ω)W1,p0(Ω)]andY=Lp(Ω)×Lp(Ω).

    Let

    V=(1+β1+ρu)v, (2.1)

    then

    v=V1+β1+ρug(u,V,ρ). (2.2)

    By substituting Eq (2.2) into Eq (1.1), one can see that the v equation in Eq (1.1) becomes

    0=vtΔ[(1+β1+ρu)v]+v(lv+cumu+v)=guut+gVVtΔVg(lg+cumu+g),

    where gu=Vρβ(1+ρu+β)2 and gV=1+ρu1+ρu+β. Hence (u,V) satisfies the following evolutional problem,

    {ut=Δu+u(kubgmu+g),guut+gVVt=ΔV+g(lg+cumu+g). (2.3)

    The monotonicity of g=g(u,V,ρ) in u and V guarantees that a one-to-one correspondence is formed by Eq (2.1) between each solution (u,v) of Eq (1.1) and each solution (u,V) of Eq (2.3). Therefore, the positive nonconstant steady state (u,v) of Eq (1.1) exists if and only if the positive nonconstant steady state (u,V) of Eq (2.3) does. Obviously, the steady state (u(x,t),V(x,t)) of Eq (2.3) satisfies the following system,

    {Δu+u(kubgmu+g)=0,ΔV+g(lg+cumu+g)=0. (2.4)

    For each h(x)C1(ˉΩ), let λ1(h) be the smallest eigenvalue of the following elliptic problem

    {Δu+h(x)u=λu,xΩ,u=0,xΩ, (2.5)

    then it is known that λ1(h) is simple and real. Moreover, λ1(h) is strictly increasing in h(x). If h(x)0, then λ1(0) can be simply denoted by λ1 with the corresponding positive eigenfunction Φ1(x) normalized by ||Φ1||L2(Ω)=1. It is noted that when the constant a>λ1, the following Dirichlet problem of elliptic equation

    {Δu(x)+u(x)(au(x))=0,xΩ,u(x)=0,xΩ

    has a unique positive solution denoted by u(x)=θa(x). Hence the stationary problem (2.4) has a semi-trivial solution (u,V)=(θk(x),0) if k>λ1.

    For convenience of our later use, we will show the nonexistence results and some priori estimates of the positive solution for Eq (2.4), which have been proved in [20].

    Lemma 2.1. (i) If kλ1,lλ1cm, then (2.4) has no positive solution.

    (ii) Assume that (u,V) and (u,v) are positive steady states of Eqs (2.3) and (1.1), respectively, then for each xΩ, it holds that

    0<u(x)k,0<v(x)V(x)(l+cm)(1+β).

    In the following, we try to obtain the positive solutions of Eq (2.4) emanating from the semi-trivial solution (θk(x),0) by regarding ρ as the bifurcation parameter.

    Define the operator F:X×R+Y by

    F(u,V,ρ)=(Δu+u(kubg(u,V,ρ)mu+g(u,V,ρ))ΔV+g(u,V,ρ)(lg(u,V,ρ)+cumu+g(u,V,ρ)))(F1(u,V,ρ)F2(u,V,ρ)). (2.6)

    Denote the Frˊechet derivative of F with respect to u and V by D(u,V)F(u,V,ρ) as follows,

    D(u,V)F(θk,0,ρ)(ϕψ)=(Δϕ+(k2θk)ϕbmgVψΔψ+(l+cm)gVψ) (2.7)

    with gV=gV(θk,0,ρ). For the sake of applying the local bifurcation theorem proposed by Crandall and Rabinowtiz [21], we firstly show that the kernel space of D(u,V)F(θk,0,ρ) is nontrivial. If (ϕ(x),ψ(x))KerD(u,V)F(θk,0,ρ), it follows from Eq (2.7) that

    {Δϕ+(k2θk)ϕb(1+ρθk)m(1+ρθk+β)ψ=0,Δψ+(l+cm)1+ρθk1+ρθk+βψ=0. (2.8)

    To obtain the nontrivial solution of the system (2.8), we need to find some ρ such that S(k,l,ρ)=0 holds for any k>λ1 and l>λ1cm, where S(k,l,ρ) is defined below.

    Lemma 2.2. Assume that λ1(h) be the smallest eigenvalue of (2.5) and λ1=λ1(0). Let

    S(k,l,ρ)λ1((l+cm)1+ρθk1+ρθk+β),k>λ1,l>λ1cm

    and

    Γ:={(k,l,ρ)R3+:S(k,l,ρ)=0},

    then the set Γ can be expressed by

    Γ:={(k,l,ρ)R3+,ρ=ρ1(k,l)fork>λ1,λ1cm<l<(1+β)λ1cm},

    where ρ=ρ1(k,l) is a positive continuous function.

    Proof. Note that

    ρ[(l+cm)1+ρθk1+ρθk+β]=(l+cm)βθk(1+ρθk+β)2<0

    and the mapping q(x)λ1(q):C(ˉΩ)R is smooth and strictly increasing. Thus, it follows that

    ρS(k,l,ρ)<0for all(k,l,ρ)(λ1,+)×(λ1cm,+)×R+. (2.9)

    With the similar proof of Lemma 6 in [20], we can prove that there exists a continuous function l(k,ρ) satisfying S(k,l,ρ)=0 and λ1cm<l<(1+β)λ1cm for any k>λ1,ρ>0. Then for any fixed k0>λ1 and l0(λ1cm,(1+β)λ1cm), there exists a unique ρ0>0 such that S(k0,l0,ρ0)=0. In virtue of the Implicit Function Theorem and Eq (2.9), it follows that there exists a small ε>0 and a unique function ρ=ρ1(k,l) with (k,l)(k0ε,k0+ε)×(l0ε,l0+ε) such that S(k,l,ρ1(k,l))=0. As k0 and l0 are arbitrary, it can be obtained that there exists a smooth function ρ=ρ1(k,l) such that S(k,l,ρ1(k,l))=0 for (k,l)(λ1,+)×(λ1cm,(1+β)λ1cm), which completes the proof.

    For ease of notation, we simplify ρ1(k,l) as ρ1. From Lemma 2.2, there exists a positive function denoted by ψ(x) which solves the following eigenvalue problem

    {Δψ+(l+cm)1+ρ1θk1+ρ1θk+βψ=0inΩ,ψ=0onΩ (2.10)

    with Ω(ψ)2dx=1. Now we give the main result in this section.

    Theorem 2.3. Assume k>λ1 and λ1<l+cm<(1+β)λ1 hold. Then there exist a small δ>0 and a smooth function ρ(s)C(0,δ] with ρ(0)=ρ1 such that the stationary problem (2.4) has a positive nonconstant solution (u(x,s),V(x,s)) bifurcating from (θk(x),0), which satisfies the following expression:

    [u(x,s)V(x,s)]=[θk(x)0]+s[ϕ(x)ψ(x)]+s[u1(x,s)V1(x,s)]  for s(0,δ),

    where

    ϕ=(Δk+2θk)1(b(1+ρ1θk)m(1+ρ1θk+β)ψ)<0

    and (u1(,s),V1(,s))C[(0,δ),X] satisfying u1(x,0)=V1(x,0)=0.

    Proof. By Lemma 2.2 and Eq (2.10), it can be deduced that

    kerD(u,V)F(θk,0,ρ1)=span{(ϕ,ψ)}. (2.11)

    Under the Dirichlet boundary condition, the operator Δk+2θk is proved to be invertible in [22]. Moreover, when ϕ is positive, so is (Δk+2θk)1ϕ. Next, we will certify codim[Range D(u,V)F(θk,0,ρ1)]=1. Suppose that (ξ,η)Range D(u,V)F(θk,0,ρ1), then there exists (ϕ,ψ)X such that

    {Δϕ+(k2θk)ϕ+b(1+ρ1θk)m(1+ρ1θk+β)ψ=ξ,Δψ+(l+cm)1+ρ1θk1+ρ1θk+βψ=η. (2.12)

    By Eq (2.10) and applying Fredholm alternative theorem, there exists a solution ˜ψ for the second equation of Eq (2.12) if and only if ηψdx=0. Substituting ˜ψ into the first equation of Eq (2.12), it has a unique solution ˜ϕ due to the invertibility of the operator Δk+2θk. Then, we can make a conclusion that codim[Range D(u,V)F(θk,0,ρ1)]=1.

    For the purpose of applying the local bifurcation argument at (ϕ,ψ,ρ)=(θk,0,ρ1), we need to certify that

    D2(u,V),ρF(θk,0,ρ1)(ϕψ)RangeD(u,V)F(θk,0,ρ1), (2.13)

    where

    D2(u,V),ρF(θk,0,ρ1)(ϕψ)=(βθkbm(1+ρθk+β)2ϕβ(l+cm)θk(1+ρθk+β)2ψ). (2.14)

    By contradiction, suppose that

    D2(u,V),ρF(θk,0,ρ1)(ϕψ)RangeD(u,V)F(θk,0,ρ1).

    Then we can find some (ˆϕ,ˆψ)X such that

    Δˆψ+(l+cm)1+ρ1θk1+ρ1θk+βˆψ=β(l+cm)θk(1+ρ1θk+β)2ψ. (2.15)

    Multiplying Eq (2.15) by ψ and integrating by parts, then by using Eq (2.10) we can obtain that

    0=(l+cm)βθk(ψ)2(1+ρ1θk+β)2dx, (2.16)

    which is impossible due to the fact that the right-hand side of Eq (2.16) is positive. Thus, this completes the proof as we have verified the transversality condition.

    Thanks to some abstract theories of stability on the basis of analytic semigroup theory (see [23]), the stability of the steady state for the Eq (2.3) can be obtained by proving the spectral stability of the steady state in W1,p(Ω)×W1,p(Ω) for p>n. In this section, we mainly investigate the distribution of spectrum. Firstly, we study the bifurcation direction which is useful for subsequently analyzing the stability of the positive steady state (u(x,s),V(x,s)).

    In the proof of Theorem 2.3, we see that

    dim{Ker[D(u,V)F(θk,0,ρ1)]}=codim{Range[D(u,V)F(θk,0,ρ1)]}=1.

    Define the adjoint operator D(u,V)F(θk,0,ρ1) of D(u,V)F(θk,0,ρ1) as follows:

    D(u,V)F(θk,0,ρ1)(ϕψ)=(Δϕ+(k2θk)ϕΔψ+(l+cm)1+ρ1θk1+ρ1θk+βψb(1+ρ1θk)m(1+ρ1θk+β)ϕ).

    Since (Δ+k2θk) is invertible, it follows that

    Range[D(u,V)F(θk,0,ρ1)]=Ker[D(u,V)F(θk,0,ρ1)]=span{(0,ψ)}whereψ>0. (3.1)

    Hence by the spectrum decomposition theorem, Eqs (2.11) and (3.1) imply that X and Y have the following direct decomposition

    X=Ker(D(u,V)F(θk,0,ρ1))XR,XR=Range(D(u,V)F(θk,0,ρ1))X,
    Y=Ker(D(u,V)F(θk,0,ρ1))Range(D(u,V)F(θk,0,ρ1))

    with a direct sum in Y.

    Lemma 3.1. For each fixed k>λ1, the bifurcating direction satisfies

    dρ(s)ds|s=0=12D2(u,V),(u,V)F(θk,0,ρ1)[(ϕψ),(ϕψ)],(0ψ)D2(u,V),ρF(θk,0,ρ1)(ϕψ),(0ψ)>0.

    Proof. The derivative dρ(s)dss=0 can be expressed in the above form by using the bifurcation formula I.6.3 in [24]. Hereafter, it remains to evaluate dρ(s)dss=0>0. From Eq (2.6), it can be calculated that

    (2uuF22uVF22VuF22VVF2)|(θk,0,ρ1)=(0(l+cm)ρ1β(1+ρ1θk+β)2(l+cm)ρ1β(1+ρ1θk+β)2(1+cm2θk)2(1+ρ1θk)2(1+ρ1θk+β)2).

    Thus,

    (ϕ,ψ)(2uuF22uVF22VuF22VVF2)(ϕψ)=2(l+cm)ρ1βψϕ(1+ρ1θk+β)2(1+cm2θk)2(1+ρ1θk)2(1+ρ1θk+β)2(ψ)2.

    By the facts that ϕ<0 and ψ>0, we have

    D2(u,V),(u,V)F(θk,0,ρ1)[(ϕψ),(ϕψ)],(0ψ)=Ω2ρ1βlm+cm(ψ)2ϕ2(1+ρ1θk)2(ψ)3(m2θk+cm2θk)(1+ρ1θk+β)2dx<0.

    Meanwhile, Eq (2.14) implies that

    D2(u,V),ρF(θk,0,ρ1)(ϕψ),(0ψ)=(l+cm)Ωβθk(ψ)2(1+ρ1θk+β)2dx>0,

    which shows that ρ(0)>0.

    For the rest of this section, we will study the stability of the nonconstant steady state (u(x,s),V(x,s),ρ(s)) near (θk,0,ρ1). Linearizing the Eq (2.3) at the steady state (u(x,s),V(x,s)), we can obtain that

    {˜ϕt=Δ˜ϕ+(k2ubgmu+g)˜ϕubgumubgm(mu+g)2˜ϕubgVmu(mu+g)2˜ψ,gu˜ϕt+gV˜ψt=Δ˜ψ+gV(l2g+cumu+g)˜ψgcugV(mu+g)2˜ψ+gu(l2g+cumu+g)˜ϕ+gcgcugu(mu+g)2˜ϕ. (3.2)

    The corresponding eigenvalue problem of the linearized Eq (3.2) with the eigenvalue σ is as follows,

    {Δϕ+[k2ubgmu+gubgumubgm(mu+g)2]ϕubgVmu(mu+g)2ψ=σϕ,Δψ+[gV(l2g+cumu+g)gcugV(mu+g)2]ψ+[gu(l2g+cumu+g)+gcgcugu(mu+g)2]ϕ=σ(guϕ+gVψ). (3.3)

    Rewrite Eq (3.3) as follows

    D(u,V)F(u(x,s),V(x,s),ρ(s))(ϕψ)=σ(10gugV)(ϕψ),    ϕ,ψW2,p0(Ω). (3.4)

    Introduce an operator H:X×R+Y by

    H(u(x,s),V(x,s),ρ(s))=(10gugV)1D(u,V)F(u(x,s),V(x,s),ρ(s)). (3.5)

    In virtue of Eqs (3.4) and (3.5), the Eq (3.3) can be transformed into the following system

    H(u(x,s),V(x,s),ρ(s))(ϕψ)=σ(ϕψ). (3.6)

    With the aim of studying the spectral stability of the steady state, we need to certify that there is no eigenvalue with nonnegative real part of the linearized operator H.

    Theorem 3.2. Suppose that k>λ1, λ1<l+cm<(1+β)λ1 and ρρ1>0 small enough. When 0<sˉδ with ˉδ>0 small enough, then the bifurcating steady state (u(x,s),V(x,s)) is locally asymptotically stable.

    Proof. According to Eq (3.5), we can deduce that

    H(θk,0,ρ1)(ϕψ)=(1001+ρ1θk1+ρ1θk+β)1((Δ+k2θk)ϕb(1+ρ1θk)m(1+ρ1θk+β)ψΔψ+(l+cm)1+ρ1θk1+ρ1θk+βψ)=(1001+ρ1θk+β1+ρ1θk)((Δ+k2θk)ϕb(1+ρ1θk)m(1+ρ1θk+β)ψΔψ+(l+cm)1+ρ1θk1+ρ1θk+βψ)=((Δ+k2θk)ϕb(1+ρ1θk)m(1+ρ1θk+β)ψ1+ρ1θk+β1+ρ1θkΔψ+(l+cm)ψ). (3.7)

    From Eqs (2.10) and (2.11), it follows that

    H(θk,0,ρ1)(ϕψ)=0. (3.8)

    The above equation shows that 0 is an eigenvalue of the operator H(θk,0,ρ1) with the corresponding eigenfunction (ϕ,ψ).

    Then, we prove that 0 is the principal eigenvalue of H(θk,0,ρ1). By contradiction, we assume that the operator H(θk,0,ρ1) has a positive eigenvalue σ1 with the corresponding eigenfunction (ϕ1,ψ1)X which satisfies

    H(θk,0,ρ1)(ϕ1ψ1)=σ1(ϕ1ψ1), (3.9)

    i.e.,

    {Δϕ1+(k2θk)ϕ1b(1+ρ1θk)m(1+ρ1θk+β)ψ1=σ1ϕ1,1+ρ1θk+β1+ρ1θkΔψ1+(l+cm)ψ1=σ1ψ1. (3.10)

    Assume ψ1=0. Then Eq (3.10) implies

    ϕ1=(Δk+2θk)1(σ1ϕ1)withσ1>0,

    which is a contradiction due to the fact that (Δk+2θk)1 is a positive operator, and hence ψ10. From the second equation of Eq (3.10), it can be shown that

    Δψ1+(l+cm)1+ρ1θk1+ρ1θk+βψ1=σ11+ρ1θk1+ρ1θk+βψ1. (3.11)

    However by Lemma 2.2 and Eq (2.10), the principal eigenvalue of the operator Δ+(l+cm)1+ρθk1+ρθk+β is zero, which contradicts with σ1>0. Thus all eigenvalues of H(θk,0,ρ1) except zero are negative.

    In the following, we investigate the distribution of spectrum for the linear operator H(u(x,s),V(x,s),ρ(s)) by means of perturbation arguments (see Corollary 1.13 in [25]). For 0<s<ˉδ with ˉδ small enough, then the linear operator H(u(x,s),V(x,s),ρ(s)) admits an eigenvalue σ(s) perturbed from zero and the corresponding function (ϕ1(x,s),ψ1(x,s))XR, where

    H(u(x,s),V(x,s),ρ(s))(ϕ(x)+ϕ1(x,s)ψ(x)+ψ1(x,s))=σ(s)(ϕ(x)+ϕ1(x,s)ψ(x)+ψ1(x,s))

    with σ(0)=0 and ϕ1(x,0)=ψ1(x,0)=0. Analogously, there exists an eigenvalue μ(ρ) perturbed from zero with the corresponding continuous differential functions (ϕ1(x,ρ),ψ1(x,ρ))XR satisfying

    H(θk,0,ρ)(ϕ(x)+ϕ1(x,ρ)ψ(x)+ψ1(x,ρ))=μ(ρ)(ϕ(x)+ϕ1(x,ρ)ψ(x)+ψ1(x,ρ)) (3.12)

    with μ(ρ1)=0 and ϕ1(x,ρ1)=ψ1(x,ρ1)=0. Differentiating Eq (3.12) with respect to ρ at ρ=ρ1, it yields that

    ρH(θk,0,ρ1)(ϕψ)+H(θk,0,ρ1)(ϕ1ψ1)=μ(ρ1)(ϕψ),

    where μ denotes dμdρ. From Eq (3.1), we have

    ρH(θk,0,ρ1)(ϕψ),(0ψ)=μ(ρ1). (3.13)

    It follows from Eq (3.7) that the left side of Eq (3.13) is equal to

    Ωβθk(1+ρ1θk)2Δψψdx=Ω[(l+cm)βθk(1+ρ1θk)(1+ρ1θk+β)](ψ)2dx>0.

    Finally, combining the above results with Eq (3.7), it indicates that when ρ<ρ1, the semi-trivial steady state (θk,0) is stable and unstable when ρρ1. According to Theorem 1.16 in [25], we have

    ˙σ(0)=˙ρ(0)μ(ρ1),

    where ˙σ(s)=dσds. Moreover by Lemma 3.1, it can be deduced that ˙σ(0)<0. This implies that σ(s)<0 for 0<s<ˉδ when ˉδ is sufficiently small, which also shows that the bifurcating steady state (u(x,s),V(x,s)) is locally asymptotically stable.

    The study of spatial patterns in the distribution of organisms is an important issue in ecology. Over the past decades, a large number of papers have been published to gain a better understanding of classical prey-dependent models and ratio-dependent predator-prey models. Extinction of one or both populations in predator-prey systems has occupied the most of the predator-prey literature. Therefore, it makes sense to study that how diffusion affects the stability of the predator-prey coexistence equilibrium, i.e., the spatial patterns, in the ratio-dependent model.

    The theoretical results in Sections 2 and 3 have shown that for some region in the parameter space of the ratio-dependent model, coexistence phenomenon can appear. In this section, we perform numerical simulations to demonstrate the spatial-temporal behaviors of the ratio-dependent predator-prey model (1.1) in one dimensional case. For the sake of convenience, we take the domain Ω=(0,1). Besides that, we choose the mesh size of space and time to be Δx=0.01 and Δt=0.01. To study the effect of cross-diffusion on the dynamics of Eq (1.1), we fix k=30>λ1=π2,l=10,m=1.1,b=10,c=3,ρ=2,β=1 and choose the initial data to be small perturbations of (θk(x),0) in simulations. The numerical simulations support our theoretical results on the existence and stability of the bifurcating solutions (see Figure 1). Biologically, the results imply that the species u,v can coexist if predators hunt and favor to stay in the region where density of prey species is high. We hope that the observations in this paper will help experimental ecologists to carry out some experimental setups that will ensure biodiversity.

    Figure 1.  The formation and evolution of the bifurcating solution (u(x,t),v(x,t)) with an initial data (u(x,0),v(x,0))=(0.16sin(πx)+0.1sin(4πx),0.1+0.1cos(4πx)).

    We are very grateful to the anonymous referees for careful reading and valuable comments, which led to an improvement of the original manuscript. The first author was supported by National Natural Science Foundation of China (No. 61902238, No. 12201426) and Shanghai Science and Technology Innovation Plan Project under Grant (No. 22692194600). The second author was supported by National Natural Science Foundation of China (No. 12201426).

    The authors declare there is no conflicts of interest.



    [1] N. Shigesada, K. Kawasaki, E. Teramoto, Spatial segregation of interacting species, J. Theor. Biol., 79 (1979), 83–99. https://doi.org/10.1016/0022-5193(79)90258-3 doi: 10.1016/0022-5193(79)90258-3
    [2] K. Kuto, Full cross-diffusion limit in the stationary Shigesada-Kawasaki-Teramoto model, Ann. Inst. Henri Poincare C, 38 (2021), 1943–1959. https://doi.org/10.1016/J.ANIHPC.2021.02.006 doi: 10.1016/J.ANIHPC.2021.02.006
    [3] D. Le, Cross diffusion systems on n spatial dimensional domains, Indiana Univ. Math. J., 51 (2002), 625–643. https://doi.org/10.1512/iumj.2002.51.2198 doi: 10.1512/iumj.2002.51.2198
    [4] Q. Li, Y. Wu, Stability analysis on a type of steady state for the SKT competition model with large cross diffusion, J. Math. Anal. Appl., 462 (2018), 1048–1072. https://doi.org/10.1016/j.jmaa.2018.01.023 doi: 10.1016/j.jmaa.2018.01.023
    [5] Q. Li, Y. Wu, Existence and instability of some nontrivial steady states for the SKT competition model with large cross diffusion, Discrete Contin. Dyn. Syst., 40 (2020), 3657–3682. https://doi.org/10.3934/dcds.2020051 doi: 10.3934/dcds.2020051
    [6] Y. Lou, W. Ni, Diffusion vs cross-diffusion: an elliptic approach, J. Differ. Equations, 154 (1999), 157–190. https://doi.org/10.1006/jdeq.1998.3559 doi: 10.1006/jdeq.1998.3559
    [7] Y. Lou, W. Ni, Y. Wu, On the global existence of a cross diffusion system, Discrete Contin. Dyn. Syst., 4 (1998), 193–203. https://doi.org/10.3934/dcds.1998.4.193 doi: 10.3934/dcds.1998.4.193
    [8] Y. Lou, W. Ni, S. Yotsutani, On a limiting system in the Lotka-Volterra competition with cross diffusion, Discrete Contin. Dyn. Syst., 10 (2004), 435–458. https://doi.org/10.3934/dcds.2004.10.435 doi: 10.3934/dcds.2004.10.435
    [9] W. Ni, Y. Wu, Q. Xu, The existence and stability of nontrivial steady states for S-K-T competition model with cross-diffusion, Discrete Contin. Dyn. Syst., 34 (2014), 5271–5298. https://doi.org/10.3934/dcds.2014.34.5271 doi: 10.3934/dcds.2014.34.5271
    [10] L. Wang, Y. Wu, Q. Xu, Instability of spiky steady states for S-K-T biological competing model with cross-diffusion, Nonlinear Anal., 159 (2017), 424–457. https://doi.org/10.1016/j.na.2017.02.026 doi: 10.1016/j.na.2017.02.026
    [11] K. Kuto, Y. Yamada, Coexistence problem for a prey-predator model with density-dependent diffusion, Nonlinear Anal. Theory Methods Appl., 71 (2009), e2223–e2232. https://doi.org/10.1016/j.na.2009.05.014 doi: 10.1016/j.na.2009.05.014
    [12] Y. Wang, W. Li, Stationary problem of a predator-prey system with nonlinear diffusion effects, Comput. Math. Appl., 70 (2015), 2102–2124. https://doi.org/10.1016/j.camwa.2015.08.033 doi: 10.1016/j.camwa.2015.08.033
    [13] H. Yuan, J. Wu, Y. Jia, H. Nie, Coexistence states of a predator-prey model with cross-diffusion, Nonlinear Anal. Real World Appl., 41 (2018), 179–203. https://doi.org/10.1016/j.nonrwa.2017.10.009 doi: 10.1016/j.nonrwa.2017.10.009
    [14] J. Cao, H. Sun, P. Hao, P. Wang, Bifurcation and turing instability for a predator-prey model with nonlinear reaction cross-diffusion, Appl. Math. Modell., 89 (2021), 1663–1677. https://doi.org/10.1016/j.apm.2020.08.030 doi: 10.1016/j.apm.2020.08.030
    [15] P. Abrams, L. Ginzburg, The nature of predation: prey dependent, ratio-dependent, or neither? Trends Ecol. Evol., 15 (2000), 337–341. https://doi.org/10.1016/S0169-5347(00)01908-X doi: 10.1016/S0169-5347(00)01908-X
    [16] Y. Kuang, E. Beretta, Global qualitative analysis of a ratio-dependent predator-prey system, J. Math. Biol., 36 (1998), 389–406. https://doi.org/10.1007/s002850050105 doi: 10.1007/s002850050105
    [17] M. Haque, Ratio-dependent predator-prey models of interacting populations, Bull. Math. Biol., 71 (2009), 430–452. https://doi.org/10.1007/s11538-008-9368-4 doi: 10.1007/s11538-008-9368-4
    [18] R. Peng, M. Wang, Qualitative analysis on a diffusive prey-predator model with ratio-dependent functional response, Sci. China, Ser. A Math., 51 (2008), 2043–2058. https://doi.org/10.1007/s11425-008-0037-8 doi: 10.1007/s11425-008-0037-8
    [19] G. Skalski, J. Gilliam, Functional responses with predator interference: viable alternatives to the Holling type Ⅱ model, Ecology, 82 (2001), 3083–3092. https://doi.org/10.1890/0012-9658(2001)082[3083:frwpiv]2.0.co;2 doi: 10.1890/0012-9658(2001)082[3083:frwpiv]2.0.co;2
    [20] N. Kumari, N. Mohan, Coexistence states of a ratio-dependent predator-prey model with nonlinear diffusion, Acta Appl. Math., 176 (2021), 11. https://doi.org/10.1007/s10440-021-00455-w doi: 10.1007/s10440-021-00455-w
    [21] M. Crandall, P. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321–340. https://doi.org/10.1016/0022-1236(71)90015-2 doi: 10.1016/0022-1236(71)90015-2
    [22] E. Dancer, On positive solutions of some pairs of differential equations, Trans. Amer. Math. Soc., 284 (1984), 729–743. https://doi.org/10.1090/S0002-9947-1984-0743741-4 doi: 10.1090/S0002-9947-1984-0743741-4
    [23] D. Henry, Geometric Theory of Semilinear Parabolic Equations, Springer-Verlag, Berlin-New York, 1981. https://doi.org/10.1007/bfb0089647
    [24] H. Kielh¨ofer, Bifurcation Theory: An Introduction with Applications to PDEs, Springer-Verlag, New York, 2004. https://doi.org/10.1002/zamm.200590030
    [25] M. Crandall, P. Rabinowitz, Bifurcation, perturbation of simple eigenvalues and linearized stability, Arch. Ration. Mech. Anal., 52 (1973), 161–180. https://doi.org/10.1007/BF00282325 doi: 10.1007/BF00282325
  • This article has been cited by:

    1. Ying Yu, Yahui Chen, You Zhou, Cross-Diffusion-Induced Turing Instability in a Two-Prey One-Predator System, 2023, 11, 2227-7390, 2411, 10.3390/math11112411
    2. Debjit Pal, Dipak Kesh, Debasis Mukherjee, Pattern dynamics in a predator–prey model with Smith growth function and prey refuge in predator poisoned environment, 2024, 92, 05779073, 366, 10.1016/j.cjph.2024.09.015
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1646) PDF downloads(92) Cited by(2)

Figures and Tables

Figures(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog