Research article Special Issues

Local bifurcation of a Ronsenzwing-MacArthur predator prey model with two prey-taxis

  • The paper investigates the steady state bifurcation analysis in a general Ronsenzwing-MacArthur predator prey model with two prey-taxis under Neumann boundary conditions. The results show that the rich dynamics in predator prey systems with two prey taxis.

    Citation: Xue Xu, Yibo Wang, Yuwen Wang. Local bifurcation of a Ronsenzwing-MacArthur predator prey model with two prey-taxis[J]. Mathematical Biosciences and Engineering, 2019, 16(4): 1786-1797. doi: 10.3934/mbe.2019086

    Related Papers:

    [1] Yong Luo . Global existence and stability of the classical solution to a density-dependent prey-predator model with indirect prey-taxis. Mathematical Biosciences and Engineering, 2021, 18(5): 6672-6699. doi: 10.3934/mbe.2021331
    [2] Wenbin Lyu . Boundedness and stabilization of a predator-prey model with attraction- repulsion taxis in all dimensions. Mathematical Biosciences and Engineering, 2022, 19(12): 13458-13482. doi: 10.3934/mbe.2022629
    [3] Dongfu Tong, Yongli Cai, Bingxian Wang, Weiming Wang . Bifurcation structure of nonconstant positive steady states for a diffusive predator-prey model. Mathematical Biosciences and Engineering, 2019, 16(5): 3988-4006. doi: 10.3934/mbe.2019197
    [4] Tingfu Feng, Leyun Wu . Global dynamics and pattern formation for predator-prey system with density-dependent motion. Mathematical Biosciences and Engineering, 2023, 20(2): 2296-2320. doi: 10.3934/mbe.2023108
    [5] Wanxiao Xu, Ping Jiang, Hongying Shu, Shanshan Tong . Modeling the fear effect in the predator-prey dynamics with an age structure in the predators. Mathematical Biosciences and Engineering, 2023, 20(7): 12625-12648. doi: 10.3934/mbe.2023562
    [6] Paulo Amorim, Bruno Telch, Luis M. Villada . A reaction-diffusion predator-prey model with pursuit, evasion, and nonlocal sensing. Mathematical Biosciences and Engineering, 2019, 16(5): 5114-5145. doi: 10.3934/mbe.2019257
    [7] Yun Kang, Sourav Kumar Sasmal, Komi Messan . A two-patch prey-predator model with predator dispersal driven by the predation strength. Mathematical Biosciences and Engineering, 2017, 14(4): 843-880. doi: 10.3934/mbe.2017046
    [8] Guangxun Sun, Binxiang Dai . Stability and bifurcation of a delayed diffusive predator-prey system with food-limited and nonlinear harvesting. Mathematical Biosciences and Engineering, 2020, 17(4): 3520-3552. doi: 10.3934/mbe.2020199
    [9] Peter A. Braza . Predator-Prey Dynamics with Disease in the Prey. Mathematical Biosciences and Engineering, 2005, 2(4): 703-717. doi: 10.3934/mbe.2005.2.703
    [10] Yue Xing, Weihua Jiang, Xun Cao . Multi-stable and spatiotemporal staggered patterns in a predator-prey model with predator-taxis and delay. Mathematical Biosciences and Engineering, 2023, 20(10): 18413-18444. doi: 10.3934/mbe.2023818
  • The paper investigates the steady state bifurcation analysis in a general Ronsenzwing-MacArthur predator prey model with two prey-taxis under Neumann boundary conditions. The results show that the rich dynamics in predator prey systems with two prey taxis.


    In addition to random diffusion of the predator and the prey, the model comprises also a prey-taxis term, which means that the spatial-temporal variations of the predator's velocity are directed by prey gradient. A chemotaxis model includes the responses of predators to the distribution of resources, that is, the foraging behavior of predators that move actively toward the higher prey density due to the prey defenses. The models with a prey-taxis term may undergo rich dynamics and generate different spatial patterns from other models without the prey-taxis.

    The predator-prey systems with prey-taxis have been widely investigated from different point of view in recent years. The existence and uniqueness of weak solutions to the two-species predator-prey model with one prey-taxis has been proved in [1]. The existence and uniqueness of weak solutions to an n×m reaction-diffusion-taxis system has been extended in [3]. The global existence of classical solutions to a three-species predator-prey model with two prey-taxis including Holling Ⅱ functional response has been investigate in [4]. Since the pattern formation of the attraction-repulsion Keller-Segel system has been studied in [7], many progress on extended models has been developed in [8,17,9,13]. The global existence and uniqueness of classical solutions to a predator-prey model with nonlinear prey-taxis has been shown in [12,6,18]. Some results about the global bifurcation of solutions for a predator-prey model with one prey-taxis is obtained in [16].

    In this paper, we consider the following general Ronsenzwing-MacArthur model with two prey-taxes under Neumann boundary conditions:

    {ut=Δu(αuv)(βuw)+u[c+Φ(v)+Ψ(w)],in(0,T)×Ω,vt=Δv+f(v)uΦ(v),in(0,T)×Ω,wt=Δw+g(w)uΨ(w),in(0,T)×Ω,un=vn=wn=0,on(0,T)×Ω,(u(0,x),v(0,x),w(0,x))=(u0(x),v0(x),w0(x))(0,0,0),inΩ, (1.1)

    where ΩRN is a bounded domain with smooth boundary; u represent the density of the predator, and v,w express the densities of two preys; c is the death rate of the predator; Φ(v)>0,Ψ(w)>0 represent the functional response of the prey; f(v),g(w) are the growth function of the prey respectively, which satisfy

    (1) f(0)=f(k1)=0,g(0)=g(k2)=0.

    (2)

    {f(v)>0,0<v<k1,f(v)<0,v>k1,and{g(w)>0,0<w<k2,g(w)<0,w>k2.

    The terms αuv and βuw are directed toward the increasing population density of v and w, respectively. In this way, the predators move in the direction of higher concentration of the prey species, where α,β indicate their prey-tactic sensitivity.

    Finally, we assume that all of the functions f,g,Φ,Ψ are of C1 class functions on R+. It follows from a standard approach (eg., cf. [2,11,12,14,16]) that the system (1.1), under these conditions, is well-posed for non-negative initial data 0(u0,v0,w0)W2,p(Ω)3 with p>N.

    The steady state solutions of the system (1.1) satisfy

    {Δu(αuv)(βuw)+u[c+Φ(v)+Ψ(w)]=0,inΩ,Δv+f(v)uΦ(v)=0,inΩ,Δw+g(w)uΨ(w)=0,inΩ,un=vn=wn=0,onΩ,(u(0,x),v(0,x),w(0,x))=(u0(x),v0(x),w0(x))(0,0,0),inΩ. (1.2)

    Using condition (2) and the positivity of both Φ and Ψ, we find that a constant vector (u,v,w) with u0 and v>0,w>0 is a solution of (1.2) iff

    {f(v)Φ(v)=u=g(w)Ψ(w),vk1,wk2,u[Φ(v)+Ψ(w)c]=0. (1.3)

    Particularly, we see that (u,v,w):=(0,k1,k2) is a constant non-negative steady state solution of (1.2) (for any c). Moreover, by suitably choosing the pairs {f,Φ} and {g,Ψ} we can make that the first two eqnations in (1.2) admit strictly positive solutions (u,v,w) satisfying u>0, 0<v<k1,0<w<k2. Finally, by choosing c=Φ(v)+Ψ(w) we find that the second equation (1.2) is satisfied.

    Below we let (u,v,w) be a constant positive steady state solution of (1.2). Take the prey taxis coefficient α as the main parameter, we analyze the solutions bifurcating from (u,v,w) for the system (1.2). The case of taking the prey taxis coefficient β as the parameter is similar. Our main bifurcation result will be given in the next §2.

    We would like to point out that there are other forms of the Ronsenzwing-MacArthur model which include the conversion rate. It remains interesting to extend our method for such models.

    Let N be the set of all positive integers, and N0:=N{0}. Let p>N, and set X:={uW2,p(Ω):un=0}, Y=Lp(Ω).

    In the sequel, we let (u,v,w) be a fixed constant positive steady state solution of (1.2) such that u>0, 0<v<k1,0<w<k2..

    By linearizing (1.2) around (u,v,w), we have an eigenvalue problem as follows:

    {ΔϕαuΔψβuΔφ+uΦ(v)ϕ+uΨ(w)φ=μϕ,inΩ,Δψ+[f(v)uΦ(v)]ψΦ(v)ϕ=μψ,inΩ,Δφ+[g(w)uΨ(w)]φΨ(w)ϕ=μφ,inΩ,ϕn=ψn=φn=0,onΩ. (2.1)

    The following result says that the eigenvalue problem (2.1) can be reduced into a sequence of matrix eigenvalue problems.

    Lemma 2.1. Let {λn} be the sequence of eigenvalues of Δ with Neumann boundary conditions satisfying 0=λ0<λ1λ2 For each n, let yn(x) be the corresponding eigenfunction for λn. Define

    An=(λnαuλn+uΦ(v)βuλn+uΨ(w)Φ(v)λn+f(v)uΦ(v)0Ψ(w)0λn+g(w)uΨ(w)). (2.2)

    Then

    1. Let μ be a complex number. Then μ is an eigenvalue of (2.1) iff there exists some nN0 such that μ is an eigenvalue of An.

    2. (u,v,w) is locally asymptotically stable with respect to (1.1) iff for every nN, all eigenvalues of An have negative real part.

    3. (u,v,w) is unstable with respect to (1.1) iff there exists an nN, such that An has at least one eigenvalue with nonnegative real part.

    Proof. The proof of the first assertion can be done by direct computations. Here we omit the details. The rest two assertions follow from the principle of the linearized stability [5,10].

    To continue, we use a direct calculation to find that the characteristic polynomial for each An has the form

    P(μ)=μ3+a2(α,λn)μ2+a1(α,λn)μ+a0(α,λn), (2.3)

    where

    a2(α,λn):=3λn+A+B,a1(α,λn):=3λ2n+[2A+2B+αuΦ(v)+βuΨ(w)]λn+[AB+uΦ(v)Φ(v)+uΨ(w)Ψ(w)]a0(α,λn):=λ3n+[A+B+αuΦ(v)+βuΨ(w)]λ2n+[AB+uΦ(v)Φ(v)+uΨ(w)Ψ(w)+βAuΨ(w)+αBuΦ(v)]λn+uΨ(w)Ψ(w)A+uΦ(v)Φ(v)B,

    where

    A:=uΦ(v)f(v),B:=uΨ(w)g(w). (2.4)

    We impose the following conditions for the given constant equilibrium (u,v,w):

    A>0,B>0,Φ(v)>0,Ψ(w)>0. (2.5)

    Under (2.5) we see that a2(α,λn)>0 for any nN. Hence, as a result of applying the RouthHurwitz criterion (see [8]), we have:

    Corollary 2.2. The following assertions are true.

    1. (u,v,w) is locally asymptotically stable for (1.1) iff for each nN there holds that (S1)a0(α,λn)>0,anda2(α,λn)a1(α,λn)a0(α,λn)>0.

    2. (u,v,w) is unstable for (1.1) iff there exists an nN such that (S2)a0(α,λn)0,ora2(α,λn)a1(α,λn)a0(α,λn).

    We now investigate the boundary between the stability and instability regimes

    a0(α,λn)=0,andT(α,λn):=a2(α,λn)a1(α,λn)a0(α,λn)=0.

    Let

    S:={(α,p)R2+:a0(α,p)=0}

    be the steady state bifurcation curve, and

    H:={(α,p)R2+:T(α,p)=0}

    be the Hopf bifurcation curve (see [14]). We study the steady state bifurcation form the given constant equilibrium (u,v,w).

    Since a0(α,p) is linear for α, we solve α from the equation a0(α,p)=0 and obtain that α=αS(p) is given by

    uΦ(v)αS(p)=1p(p+B)[p3+(A+B+βuΨ(w))p2+(AB+uΦ(v)Φ(v)+uΨ(w)Ψ(w)+βAuΨ(w))p+uΨ(w)Ψ(w)A+uΦ(v)Φ(v)B]=p+A+uΦ(v)Φ(v)p+β(p+A)uΨ(w)p+B+uΨ(w)Ψ(w)p+B+AuΨ(w)Ψ(w)p(p+B). (2.6)

    Correspondingly, we let αH(p) be the solution of T(α,p)=0, i.e., αH(p) is the graph of function about H.

    It is observed that the function αS(p) has the following properties:

    Lemma 2.3. Assume (2.5). Let αS(p) be defined by (2.6). If p>0 is a critical point of αS(p), then p is a local maximum point. Moreover, limpαS(p)=.

    Proof. Differentiating (2.6), we obtain

    uΦ(v)αS(p)=1uΦ(v)Φ(v)p2+βuΨ(w)(BA)(p+B)2uΨ(w)Ψ(w)(p+B)2AuΨ(w)Ψ(w)p(p+B)2AuΨ(w)Ψ(w)p2(p+B).

    Assume p>0 to be a critical point of αS, i.e., αS(p)=0. Then we have that

    βuΨ(w)(AB)(p+B)2+uΨ(w)Ψ(w)(p+B)2=1uΦ(v)Φ(v)p2AuΨ(w)Ψ(w)p(p+B)2AuΨ(w)Ψ(w)p2(p+B)

    and

    uΦ(v)αS(p)=2uΦ(v)Φ(v)p3+2βuΨ(w)(AB)(p+B)3+2uΨ(w)Ψ(w)(p+B)3+2AuΨ(w)Ψ(w)p2(p+B)2+2AuΨ(w)Ψ(w)p(p+B)3+2AuΨ(w)Ψ(w)p3(p+B)=2uΦ(v)Φ(v)p3+2p+B(1uΦ(v)Φ(v)p2)+2AuΨ(w)Ψ(w)p3(p+B)=1(p+B)[uΦ(v)Φ(v)p2+2+2u(AΨ(w)Ψ(w)+BΦ(v)Φ(v))p3].

    We find by (2.5) that all trems in the above bracket are positive, and uΦ(v)>0. This implies that αS(p)<0 and thus the critical point p>0 is a local maximum point.

    Clearly, we see from (2.5) that limpαS(p)=, since u>0 and Φ(v)>0.

    In this section, we inverstigate the global steady state bifurcation from (u,v,w) near α=αS. According to Lemma 2.1, we have the following result.

    Proposition 2.4. Let {λn} be the sequence of eigenvalues of Δ with Neumann boundary conditions, such that 0=λ0<λ1λ2, let yn(x) be the eigenfunction corresponding to λn (nN). Let αS(p) be defined by (2.6). For nN we define

    αSn=αS(λn). (2.7)

    Then the eigenvalue problem (2.1) has an eigenvalue μ=0 if and only if α=αSn for some nN, and the corresponding eigenfunction is Vnyn, where Vn satisfies AnVn=0 with An defined as in (2.2).

    We recall the following global bifurcation theorem (see [8,11]):

    Lemma 2.5. Let V be an open connected subset of R×X and (λ0,u0)V, and let F be a continuously differentiable mapping from V into Y. Suppose that

    (1) F(λ,u0)=0 for (λ,u0)V,

    (2) the partial derivative DλuF(λ,u) exists and is continuous in (λ,u) near (λ0,u0),

    (3) DuF(λ0,u0) is a Fredholm operator with index 0, and dimN(DuF(λ0,u0))=1,

    (4) Dλ(DuF(λ0,u0))[w0]R(DuF(λ0,u0)), where w0X spans N(DuF(λ0,u0)).

    Let Z be any complement of span{w0} in X. Then there exists an open interval I1=(ϵ,ϵ) and continuous functions λ:I1R,ψ:I1Z, such that λ(0)=λ0,ψ(0)=0, and, if u(s)=u0+sw0+sψ(s)for sI1, then F(λ(s),u(s))=0. Moreover, F1({0}) near (λ0,u0) consists precisely of the cures u=u0 and Γ={(λ(s),u(s)):sI1}. If in addition, DuF(λ,u) is a Fredholm operator for all (λ,u)V, then the curve Γ is contained in C, which is a connected component of ˉS where S={(λ,u)V:F(λ,u)=0,uu0}; and either C is not compact in V, or C contains a point (λ,u0) with λλ0.

    We obtain the result for the global bifurcation of steady state solutions in predator prey taxis system (1.1) as follows.

    Theorem 2.6. Assume (2.5) as well as β>0. Let αSn be defined as (2.7). Moreover, we assume that the following conditions (A1)-(A2) hold true:

    (A1)For some jN, λj is a simple eigenvalue of Δ in Ω with Neumann boundary conditions, and the corresponding eigenfunction is yj(x).

    (A2)For any nN, H(αSj,λn)0, and if nj, then αSjαSn.

    Then there hold following assertions:

    1. The system (1.2) has a unique one-parameter family Γj={(ˆUj(s),ˆαj(s)):ε<s<ε} of nontrivial solutions near (u,v,w,α)=(u,v,w,αSj). More precisely, there exists ε>0 and C function s(ˆUj(s),ˆαj(s)) from s(ε,ε) to X3×R satisfying

    (ˆUj(0),ˆαj(0))=((u,v,w),αSj),

    and

    ˆUj(s)=(u,v,w)+syj(x)(λj+uΦ(v)f(v),Φ(v),Ψ(w)(λj+uΦ(v)f(v))λj+uΨ(w)g(w))+s(h1,j(s),h2,j(s),h3,j(s)),

    such that h1,j(0)=h2,j(0)=h3,j(0)=0;

    2. The set Γj is a subset of a connected component Cj of ˉS, where

    S={(u,v,w,α)X3×R:(u,v,w,α)is a nontrivial positive equilibrium of(1.2)},

    and either Cjcontains another point (u,v,w,αSk) with αSkαSj or Cj is unbounded.

    Proof. We define a mapping F:X3×RY0×Y2 by

    F(u,v,w,α)=(Δu(αuv)(βuw)+u(c+Φ(v)+Ψ(w))Δv+f(v)uΦ(v)Δw+g(w)uΨ(w)).

    We have that F(u,v,w,αSj)=0, and F is continuously differentiable. We will prove our result by applying Lemma 2.5 to F.

    To this end, we must check that F satisfies all requirements of Lemma 2.5. It will be completed in several steps.

    (1) For each U=(u,v,w) the derivative FU(u,v,w,αSj) is a Fredholm operator with index zero (see [15]), and the kernel space N(FU(u,v,w,αSj)) is one-dimensional.

    It left to show N(FU(u,v,w,αSj)){0}. We note that

    FU(u,v,w,αSj)[ϕ,ψ,φ]=(Δϕ(αSjuψ)(βuφ)+uΦ(v)ψ+uΨ(w)φΔψAψΦ(v)ϕΔφBφΨ(w)ϕ).

    Here A,B are given by (2.4). Let (ϕ,ψ,φ)(0)FU(u,v,w,αSj), then from Lemma 2.1 there exists jN, such that 0 is an eigenvalue of Aj, and the corresponding eigenvector is

    (aj,bj,cj)yj=(λj+A,αuλj+uΦ(v),(αuλj+uΦ(v))(λj+A)λj+B)yj,

    According to condition (A1), the eigenvector is unique up to a constant multiple. Hence we have

    N(FU(u,v,w,αSj))=span{(aj,bj,cj)yj}

    that is dim(N(FU(u,v,w,αSj)))=1.

    (2) FαU(u,v,w,αSj)[(aj,bj,cj)yj]R(FU(u,v,w,αSj)).

    Define

    R(FU(u,v,w,αSj))={(h1,h2,h3,r)Y0×Y2×R:Ω(¯ajh1+¯bjh2+¯cjh3)yjdx=0}, (2.8)

    where (¯aj,¯bj,¯cj) is a non-zero eigenvector fo the eigenvalue μ=0 of ATn, ATn the transpose of An and (¯aj,¯bj,¯cj):=(aj,bj,cj)yj. Furthermore, if (h1,h2,h3,r)R(FU(u,v,w,αSj)), then there exists (ϕ1,ψ1,φ1)X3, such that

    FU(u,v,w,αSj)[ϕ1,ψ1,φ1]=(h1,h2,h3,r).

    Define

    L[ϕ,ψ,φ]=(ΔϕαSjuΔψβuΔφ+uΦ(v)ψ+uΨ(w)φΔψAψΦ(v)ϕΔφBφΨ(w)ϕ)

    and its adjoint operator

    L[ϕ,ψ,φ]=(ΔϕΦ(v)ψΨ(w)φΔψαSjuΔϕ+uΦ(v)ϕAψΔφβuΔϕ+uΨ(w)ϕBφ).

    Then

    (h1,h2,h3),(¯aj,¯bj,¯cj)yj=L[(ϕ1,ψ1,φ1)],(¯aj,¯bj,¯cj)yj=(ϕ1,ψ1,φ1),L[(¯aj,¯bj,¯cj)yj]=(ϕ1,ψ1,φ1),An[(¯aj,¯bj,¯cj)yj],

    where , is the inner product in [L2(Ω)]3. If

    (h1,h2,h3,r)R(FU(u,v,w,αSj)),

    then

    Ω(¯ajh1+¯bjh2+¯cjh3)yjdx=0. (2.9)

    Since (2.9) defines a codimension-1 set in Y0×Y2×R, we get that

    codimR(FU(u,v,w,αSj))=dimN(FU(u,v,w,αSj))=1.

    Therefore, R(FU(u,v,w,αSj)) is given by (2.8).

    Note that

    FαU(u,v,w,αSj)[(aj,bj,cj)yj]=(ubjΔyj,0,0,0)=(uΦ(v)λjyj,0,0,0).

    Therefore,

    Ω(¯ajh1+¯bjh2+¯cjh3)yjdx=Ω(λj+A)uΦ(v)λjyjdx>0.

    Here we have used the condition A>0 from (2.5). Thus FαU(u,v,w,αSj)[(aj,bj,cj)yj]R(FU(u,v,w,αSj)).

    Finally, we can apply the argument in Lemma 2.3 of [15] to obtain that the operator FU(u,v,w,αSj) is a Fredholm operator with index zero for any (u,v,w,λ)X3×R. Hence, all conditions in Lemma 2.5 are satisfied, and we get that solutions bifurcating from (u,v,w,αSj) are on a connected component Cj of the set of nontrivial solutions of (1.2). We see that all solutions on Cj are positive. This is apparently true for solutions near the bifurcation point (u,v,w,αSj) by u>0,v>0,w>0. From the equation of v and w in (1.2), we know that if u is nonnegative then there exists (u,v,w,λ)Cj, such that v(x)>0,w(x)>0 in ˉΩ, and when xˉΩ we have u(x)=0. But from the equation of u in (1.2), which is linear about u, we see that this is a contradiction by the strong maximum principle. Thus all solution are positive on Cj.

    Some numerical simulations of (1.1) are shown in this section, here the functional response Φ(v)=m1va1+v, Ψ(w)=m2va2+v are taken as Holling type Ⅱ, and f(v)=v(k1v) and g(w)=w(k2w) are taken as Logistic growth.

    For simplicity, we take a1=a2=a, k1=k2=k and m1=m2=m. Then the Eqn. (1.3) for a solution (u,v,w) with u>0 becomes

    {(vk)(v+a)/m=u>0,w=v(0,k),v/(v+a)=c/(2m). (2.10)

    Thus, we assume that

    0<˜v:=ac2mc<k. (2.11)

    Then, under condition (2.11), we obtain solution of (2.10) and thus constant steady state (u,v,w) of (1.2) as follows:

    v=w=˜v,u=(˜vk)(˜v+a)>0.

    In the following, we fix

    c=0.5,a=0.56,k=2.

    Then m satisfies condition (2.11) iff

    m>0.32. (2.12)

    Moreover, we take Ω=(0,30π) (one-dimensional space).

    1. We take m1=0.35,m2=0.35 so that m:=0.35 satisfies condition (2.12). Then the constant steady state (u,v,w) is locally asymptotically stable for α=β=0, see Figure 1, where the initial value is (u0,v0,w0)=(3.3+0.01sin(10x);1.4+0.02sin(x);1.4+0.02sin(x)). From Theorem 2.6, the constant steady state (u,v,w) becomes unstable where a steady state bifurcates from it for α=100,β=1, see Figure 2.

    Figure 1.  some non-negative solutions of (1.1) converge to the positive constant steady states (u,v,w).
    Figure 2.  a non-negative solutions of (1.1) converge to some non-constant steady state.

    2. We further find that the taxis coefficient α and β play a stability role for (1.1). In fact, we take m1=0.43,m2=0.43 so that m:=0.43 satisfies condition (2.12). Then there is a periodic solution bifurcating from the positive constant steady state (u,v,w) for α=β=0, see Figure 3, where the initial value is (u0,v0,w0)=(3.9,0.75,0.75). However, the positive large taxis coefficients α=1×105 and β=1×104 hold back the oscillation and make (u,v,w) more stable, see Figure 4, where the initial value is (u0,v0,w0)=(2,1,1).

    Figure 3.  the periodic solutions bifurcating from (u,v,w) of (1.1).
    Figure 4.  large taxis coefficients α and β result in the convergence to (u,v,w).

    We might expect such an ecosystem (1.1) to exhibit a rich dynamical interplay among the three species. In this paper, we show that this is indeed the case, see Figure 1Figure 4.

    We investigate the steady state bifurcation analysis in a general Ronsenzwing-MacArthur predator prey model with two prey-taxis under Neumann boundary conditions. Comparing with the dynamic analysis of reaction diffusion predator prey systems without taxis, the bifurcation analysis become more complicated and the bifurcation results in this paper cover most of reaction diffusion taxis predator prey systems. The results show that the rich dynamics in predator prey systems with two prey taxis and we will study other properties introduced by two taxis term in the future.

    We deeply thank three anonymous reviewers for their kind helps and valuable comments. We thank also Prof. Sen-Zhong Huang for valuable discussions. The work partially is supported by grants from NSFC(No. 11601105 and No. 11471091), NSFH(No. A2017007), GBD1317047, and HUDF2019101.

    The authors declare there is no conflict of interest.



    [1] B. Ainseba, M. Bendahmane and A. Noussair, A reaction-diffusion system modeling predator-prey with prey-taxis, Nonl. Anal.: Real World Applications, 9 (2008) 5, 2086–2105.
    [2] H. Amann, Dynamic theory of quasilinear parabolic equations II, Reaction-diffusion systems, Diff. Int. Eqns., 3 (1990), 13–75.
    [3] M. Bendahmane, Analysis of a reaction-diffusion system modeling predator-prey with prey-taxis, Network and Heterogeneous Media, 3 (2008), 863–879.
    [4] L. L. Chen and G. H. Zhang, Global existence of classical solutions to a three-species predatorprey model with two prey-taxis, J. Appl. Math., 10(2012), 1155–1167.
    [5] A. K. Drangeid, The principle of linearized stability for quasilinear parabolic evolution equations, Nonlinear Analysis, 13 (1989), 1091–1113.
    [6] H. Y. Jin and Z. A. Wang. Global stability of prey-taxis systems, J. Diff. Eqns., 3(2017), 1257– 1290.
    [7] E. F. Keller and L. A. Segel. Initiation of slime mold aggregation viewed as an instability, J. Theoret. Biol., 26(3) (1970) 399–415.
    [8] P. Liu, J. P. Shi and Z. A. Wang, Pattern formation of attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2597–2625.
    [9] D. M. Liu and Y. S. Tao, Global boundedness in a fully parabolic attraction-repulsion chemotaxis model, Math. Methods Appl. Sci., 38 (2015), 2537–2546.
    [10] G. Simonett, Center manifolds for quasilinear reaction-diffusion systems, Diff. Integr. Eqns., 8 (1995), 753–796.
    [11] J. P. Shi and X. F. Wang, On global bifurcation for quasilinear elliptic systems on bounded domains, J. Diff. Eqns., 246 (2009), 2788–2812.
    [12] Y. Tao, Global existence of classical solutions to a predator-prey model with nonlinear prey-taxis, Nonl. Anal.: Real World Applications, 11 (2010), 2056–2064.
    [13] Y. S. Tao. Global dynamics in a higher-dimensional repulsion chemotaxis model with nonlinear sensitivity, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2705–2722.
    [14] J. F. Wang, J. P. Shi and J. J. Wei, Dynamics and pattern formation in a diffusion predator-prey system with strong Allee effect in prey, J. Diff. Eqns., 251 (2011), 1276–1304.
    [15] X. F. Wang and Q. Xu, Spiky and transition layer steady states of chemotaxis systems via global bifurcation and Helly's compactness, J. Math. Biol., 66 (2013), 1241–1266.
    [16] X. L. Wang, W. D. Wang and G. H. Zhang, Global bifurcation of solutions for a predator-prey model with prey-taxis, Math. Meth. Appl. Sci., 3 (2014), 431–443.
    [17] M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model, J. Diff. Eqns., 248 (2010) (12), 2889–2905.
    [18] S. N. Wu, J. P. Shi, and B. Y. Wu. Global existence of solutions and uniform persistence of a diffusive predator-prey model with prey-taxis, J. Diff. Eqns., 260 (2016), 5847–5874.
  • This article has been cited by:

    1. Gurusamy Arumugam, Global existence and stability of three species predator-prey system with prey-taxis, 2023, 20, 1551-0018, 8448, 10.3934/mbe.2023371
  • Reader Comments
  • © 2019 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(4550) PDF downloads(610) Cited by(1)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog