
Citation: Dongfu Tong, Yongli Cai, Bingxian Wang, Weiming Wang. Bifurcation structure of nonconstant positive steady states for a diffusive predator-prey model[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 3988-4006. doi: 10.3934/mbe.2019197
[1] | Yuxuan Zhang, Xinmiao Rong, Jimin Zhang . A diffusive predator-prey system with prey refuge and predator cannibalism. Mathematical Biosciences and Engineering, 2019, 16(3): 1445-1470. doi: 10.3934/mbe.2019070 |
[2] | Tingting Ma, Xinzhu Meng . Global analysis and Hopf-bifurcation in a cross-diffusion prey-predator system with fear effect and predator cannibalism. Mathematical Biosciences and Engineering, 2022, 19(6): 6040-6071. doi: 10.3934/mbe.2022282 |
[3] | 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 |
[4] | Jun Zhou . Bifurcation analysis of a diffusive plant-wrack model with tide effect on the wrack. Mathematical Biosciences and Engineering, 2016, 13(4): 857-885. doi: 10.3934/mbe.2016021 |
[5] | Yongli Cai, Malay Banerjee, Yun Kang, Weiming Wang . Spatiotemporal complexity in a predator--prey model with weak Allee effects. Mathematical Biosciences and Engineering, 2014, 11(6): 1247-1274. doi: 10.3934/mbe.2014.11.1247 |
[6] | 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 |
[7] | Zuolin Shen, Junjie Wei . Hopf bifurcation analysis in a diffusive predator-prey system with delay and surplus killing effect. Mathematical Biosciences and Engineering, 2018, 15(3): 693-715. doi: 10.3934/mbe.2018031 |
[8] | Mingzhu Qu, Chunrui Zhang, Xingjian Wang . Analysis of dynamic properties on forest restoration-population pressure model. Mathematical Biosciences and Engineering, 2020, 17(4): 3567-3581. doi: 10.3934/mbe.2020201 |
[9] | Xue Xu, Yibo Wang, Yuwen Wang . Local bifurcation of a Ronsenzwing-MacArthur predator prey model with two prey-taxis. Mathematical Biosciences and Engineering, 2019, 16(4): 1786-1797. doi: 10.3934/mbe.2019086 |
[10] | Xiaoling Li, Guangping Hu, Xianpei Li, Zhaosheng Feng . Positive steady states of a ratio-dependent predator-prey system with cross-diffusion. Mathematical Biosciences and Engineering, 2019, 16(6): 6753-6768. doi: 10.3934/mbe.2019337 |
Understanding the mechanisms by which patterns are created in the living system poses one of the most challenging problems in developmental biology [1], since the pioneer work of Turing [2]. Turing's revolutionary idea was that passive diffusion could interact with the chemical reaction in such a way that even if the reaction by itself has no symmetry-breaking capabilities, diffusion can destabilize the symmetry so that the system with diffusion can have them [3,4,5,6].
There has been considerable interest in investigating the pattern formation of population system by taking into account the effect of diffusion [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,26,27,28,29,30,31,32,33]. These investigations have revealed that spatial inhomogeneities like the inhomogeneous distribution of nutrients as well as interactions on spatial scales like migration play an important role in specializing and stabilizing population levels.
In a recent analytic approach by Shi, Li and Lin [17], the following reaction-diffusion predator-prey model is considered:
{ut−d1uxx=u(1−u)−kuva+u+mv,x∈(0,π),t>0,vt−d2vxx=v(δ−βvu),x∈(0,π),t>0,ux=0,vx=0,x=0,π,t>0,u(x,0)=u0(x)≥0,v(x,0)=v0(x)≥0,x∈(0,π), | (1.1) |
where u:=u(x,t) and v:=v(x,t) represent the densities of the prey and predator, respectively. The parameter k>0 is the maximum consumption rate, a a saturation constant, m a predator interference parameter (m<0 the case where predators benefit from cofeeding [18]), δ the intrinsic growth rate of prey, and β the numbers of prey required to support one predator at equilibrium when v equals to u/β [34]. d1,d2 are the diffusion coefficients of u and v, respectively.
In [17], the authors show that, model (1.1) has a boundary steady state (1,0) which is unstable and a unique positive steady state
E∗=(u∗,v∗)=(β(1−a)+δ(m−k)+√(aβ−β−mδ+kδ)2+4aβ(β+mδ)2(β+mδ),δβu∗). |
And the authors investigated the qualitative properties, including the global attractor, persistence property under the condition of m>k, local and global stability of E∗, and established the existence and nonexistence of nonconstant positive steady states of model (1.1). In this paper, we always assume that m>k.
And there naturally comes a question: What is the structure of nonconstant positive steady states of model (1.1)?
Thus we will investigate the steady state problem corresponding to model (1.1)
{−d1uxx=u(1−u)−kuva+u+mv,x∈(0,π),−d2vxx=v(δ−βvu),x∈(0,π),ux=vx=0,x=0,π. | (1.2) |
It is our purpose in this paper to make a better description for the structure of the set of nonconstant positive steady states of model (1.1). The rest of this article is organized as follows: In Section 2, we prepare some preliminaries and give the Turing instability in details. In Section 3, we give the local and global bifurcation structure of nonconstant positive steady states.
It's well known that the Turing instability refers to "diffusion driven instability", i.e., the stability of the positive equilibrium E∗=(u∗,v∗) changing from stable for the ordinary differential equations (ODE) dynamics (i.e., d1=d2=0 in model (1.1)), to unstable, for the partial differential equations (PDE) dynamics (1.1) [1,14]. The occurrence of Turing pattern is caused by the existence of nonconstant positive steady states of model (1.1) as a result of diffusion. In this section, we mainly discuss Turing instability.
Before proceeding, we recall the following Neumann eigenvalue problem
{−φxx=λφ,x∈(0,π),φx=0,x=0,π. | (2.1) |
It is well known that (2.1) has a sequence of simple eigenvalues and associated eigenfunctions are explicitly given by
λi=i2,φi(x)={1√π,i=0,√2πcosix,i>0, |
where i=0,1,2,⋅⋅⋅. Let Y=C2((0,π))×C2((0,π)) be the Hilbert space, and
X={(u,v)∣u,v∈C2((0,π)),ux=vx=0,x=0,π}. |
Let us first recall that the ODE model corresponding to the PDE model (1.1):
{dudt=u(1−u)−kuva+u+mv:=f(u,v),dvdt=v(δ−βvu):=g(u,v). | (2.2) |
Following [17], the Jacobian of model (2.2) around E∗=(u∗,v∗) is given by
J=(a11a12a21a22), | (2.3) |
where
a11=u∗(β(1−a)−(mδ+2β)u∗)(mδ+β)u∗+aβ,a12=−kβ2u∗(a+u∗)(mδ+β)u∗+aβ,a21=δ2β,a22=−δ. | (2.4) |
The characteristic equation of J is
η2−Tr(J)η+det(J)=0, |
where
Tr(J)=−βu∗2+δ(δm+β+m−k)u∗+αβ(δ+1)(mδ+β)u∗+aβ,det(J)=δu∗((δm+β)(β(1+a)+δ(m−k)u∗+aβ(aβ+δk+δm+β)))((mδ+β)u∗+aβ)2>0. | (2.5) |
It follows from the assumption m>k that Tr(J)<0. Thus the positive equilibrium E∗ of the ODE model (2.2) is locally stable.
Choosing d2 as the bifurcation parameter, we have the following linearized operator of model (1.2) evaluated at E∗:
L(d2)=(d1∂2∂x2+a11a12a21d2∂2∂x2+a22). |
It is easy to see that the eigenvalues of L(d2) are given by those if the following operator Li(d2) (see, e.g., [16,17,36]):
Li=(−d1i2+a11a12a21−d2i2+a22), |
whose characteristic equation is
ξ2−ξTi+Qi=0,i=0,1,2,⋯ |
where
Ti=−(d1+d2)i2+Tr(J)<0,Qi=i2(d1i2−a11)(d2−d1δi2+det(J)i2(a11−d1i2)). | (2.6) |
For simplicity, we define
θ:=(1−a)β−(mδ+2β)u∗. | (2.7) |
If θ>0 and
d1<a11, | (2.8) |
then what we define as i0:=i0(k,a,m,δ,β) is the largest positive integer such that d1i2<a11 with i≤i0.
Clearly, if (2.8) is satisfied, then 1≤i0≤∞. In this case, we denote
¯d2=min0≤i≤i0di2,wheredi2:=d1δi2+det(J)i2(a11−d1i2). | (2.9) |
Therefore we can obtain the local stability of E∗=(u∗,v∗) of model (1.1) as follows:
Theorem 2.1. For model (1.1),
(ⅰ) If θ<0 holds, then E∗ is locally asymptotically stable.
(ⅱ) If θ>0, we have
(ⅰ-1) if d1<a11 and 0<d2<¯d2 hold, then E∗ is locally asymptotically stable.
(ⅱ-2) If d1≤a11,d2>¯d2 hold, then E∗ is Turing unstable.
Example 2.2. As an example, we take the parameters in model (1.1) as:
a=0.035,m=0.05,k=0.8,β=1.5,δ=2,d1=0.01 |
Easy to know there is a unique positive equilibrium E∗=(u∗,v∗)=(0.16548,0.22064). From Theorem 2.1, we can know that if d2>¯d2=0.19 holds, E∗ is Turing unstable, and model (1.1) exhibits Turing pattern. In Figure 1, we show the numerical results of model (1.1) with different values of d2. Figure 1a shows Turing pattern with d2=0.21>¯d2, and one can see that the solutions of (1.1) are not dependent on time t but space x. In other words, model (1.1) in this case has nonconstant positive steady states as a result of diffusion. Figure 1b gives the stable behavior of model (1.1) with d2=0.15<¯d2.
In this section, we will focus on the local and global bifurcation structure of the nonconstant positive steady states for model (1.2).
If (u,v)=(u(x),v(x)) is a positive solution to model (1.2), then it is proved in [17] and the maximum principle [35] that
1−km<u(x)<1,αβ(1−km)<v(x)<αβ,x∈Ω. | (3.1) |
We translate (u∗,v∗) to the origin by the translation (˜u,˜v)=(u−u∗,v−v∗). For convenience, we still denote ˜u,˜v by u,v, respectively, and then we can obtain the following system
{−d1uxx=(u+u∗)(1−(u+u∗))−k(u+u∗)(v+v∗)a+(u+u∗)+m(v+v∗)=:f(u+u∗,v+v∗),x∈(0,π),−d2vxx=(v+v∗)(δ−β(v+v∗)(u+u∗))=:g(u+u∗,v+v∗),x∈(0,π),ux=vx=0,x=0,π. | (3.2) |
We study the local structure of nonconstant positive solutions for the new system (3.2). In brief, by regarding d2 as bifurcation parameter, we verify the existence of positive solutions bifurcation from (d2,0,0). In this section, we assume that θ>0.
With the help of a priori estimate (3.1), let
X={(u,v)∈C2(0,π)×C2(0,π):ux=vx=0,x=0,π} |
and Y=Lp(0,π)×Lp(0,π). Define the map F:(0,∞)×X→Y by
F(d2,u,v)=(d1∂2u∂x2+f(u+u∗,v+v∗)d2∂2v∂x2+g(u+u∗,v+v∗)). |
Then the solutions of the boundary problem (3.2) are exactly zero of the map F(d2,u,v). Note that (0,0) is the unique constant solution of (3.2), then we have F(d2,0,0)=0. The Frˊechet derivative of F(d2,u,v) with respect to (u,v) at (0,0) can be given by
L1(d2)=F(u,v)(d2,0,0)=(d1∂2∂x2+a11a12a21d2∂2∂x2+a22), |
where a11,a12,a21 and a22 are given in (2.4).
Theorem 3.1. Let
d2=dj2:=d1δλj+det(J)λj(a11−d1λj),λj=j2for1≤j≤i0. | (3.3) |
(i) Suppose that dj2≠di2 for any integer i≠j. Then (dj2,0,0) is a bifurcation point of F(d2,u,v)=0. Moreover, there is a one-parameter family of non-trivial solutions Γj(s)=(d2(s),u(s),v(s)) of F(d2,u,v)=0 for |s| sufficiently small, where d2(s),u(s),v(s) are continuous functions, (u(0),v(0))=(0,0),d2(0)=dj2 and u(s)=sφj+o(s),v(s)=sbjφj+o(s),bj=a11−d1λja12>0. The zero set of F(d2,u,v) consists of two curves (d2,0,0) and Γj(s) in a neighborhood of the bifurcation point (dj2,0,0).
(ii) Suppose that there exists a positive integer i(≠j) such that di2=dj2≜ˆd. Let
bi=a11−d1λia12,b∗i=d1λi−a11a21,Φi=(1bi)φi, | (3.4) |
X2={(u,v)∈Y:∫π0(u+biv)φidx=∫π0(u+bjv)φjdx=0}. | (3.5) |
If 1+b∗i1+bib∗i≠0,1+b∗j1+bjb∗j≠0 and j=2i(resp.i=2j), then (ˆd,0,0) is a bifurcation point of F(d2,u,v)=0. Moreover, there exists a curve of nonconstant solutions (d2(ω),s(ω)(cosωΦi+sinωΦj)+W(ω)) of F(d2,u,v)=0 for ∣ω−ω0∣ sufficiently small, where d2(ω),s(ω), and W(ω) are continuously differentiable functions with respect to ω,W(ω)∈X2 and satisfy d2(ω0)=ˆd,s(ω0)=0,W(ω0)=0. Here ω0 is any constant satisfying
cosω0≠0andc2c4bjλjsin2ω0≠c1c3biλicos2ω0, | (3.6) |
(resp.sinω0≠0andc2c5bjλjsin2ω0≠c1c6biλicos2ω0), | (3.7) |
where
c1=b∗i1+bib∗i,c2=b∗j1+bjb∗j,c3=√12πA1+B1b∗j1+bjb∗j,c4=√12πA2+B2b∗i1+bib∗i,c5=√12πA3+B3b∗i1+bib∗i,c6=√12πA2+B2b∗j1+bjb∗j, |
and
A1=(−1+kv∗(mv∗+a)(a+u∗+mv∗)3)−k(amv∗+2mu∗v∗+a2+au∗)(a+u∗+mv∗)3bi+ku∗m(a+u∗)(a+u∗+mv∗)3b2i,A2=2(−1+kv∗(mv∗+a)(a+u∗+mv∗)3)−k(amv∗+2mu∗v∗+a2+au∗)(a+u∗+mv∗)3(bi+bj)+ku∗m(a+u∗)(a+u∗+mv∗)3bibj,A3=(−1+kv∗(mv∗+a)(a+u∗+mv∗)3)−k(amv∗+2mu∗v∗+a2+au∗)(a+u∗+mv∗)3bj+ku∗m(a+u∗)(a+u∗+mv∗)3b2j,B1=−δ2βu∗+2δu∗bi−βu∗b2i,B2=−2δ2βu∗+2δu∗(bi+bj)−βu∗bibj,B3=−δ2βu∗+2δu∗bj−βu∗b2j. | (3.8) |
Proof. (ⅰ) By using the Crandall-Rabinowitz bifurcation theorem [37], we know that (dj2;0,0) is a bifurcation point provided that:
(a) the partial derivatives Fd2,F(u,v), and Fd2,(u,v) exist and are continuous;
(b) kerF(u,v)(dj2,0,0) and codimIm(F(u,v)(dj2,0,0)) are one-dimensional (here Im: image);
(c) let kerF(u,v)(dj2,0,0)=span{Φj}, then Fd2(u,v)(dj2,0,0)Φj∉Im(F(u,v)(dj2,0,0)).
It suffices to verify conditions (a)-(c) above. When d2=dj2, the operator L1(d2) is given by
L1(dj2)=F(u,v)(d2,0,0)=(d1∂2∂x2+a11a12a21dj2∂2∂x2+a22), |
It is clear that the linear operators F(u,v),Fd2,(u,v) and Fd2 are continuous. The condition (a) is verified.
Suppose Φi=(ˉϕ,ˉψ)⊤∈kerL1, and write ˉϕ=Σˉaiφi,ˉψ=Σˉbiφi. Then
∞∑i=0ˉBi(ˉaiˉbi)φi=0,whereˉBi=(a11−d1λia12a21a22−d2λi). | (3.9) |
By a simple calculation, we have detˉBi≠0, when i≠j. Hence if and only if i=j,
detˉBi=0⇔d2=di2=d1δλi−det(J)λi(a11−d1λi), |
taking d2=dj2 implies that
kerL1=span{Φj},Φj=(1bj)φj. |
Consider the adjoint operator
L∗1=(d1∂2∂x2+a11a21a12d2∂2∂x2+a22). |
In the same way as above we obtain
kerL∗1=span{Φ∗j},Φ∗j=(1b∗j)φj. |
Since Im(L1)=ker(L∗1)⊥ (here ⊥: complementary set), thus
codim(Im(L1))=dim(ker(L∗1))=1. |
Condition (b) is also verified.
Finally, since
F(u,v)(dj2,0,0)Φj=(000∂2∂x2)Φj=(0−λjbjφj), |
and
⟨Fd2,(u,v)(dj2,0,0)Φj,Φ∗j⟩Y=⟨−λjbjφj,b∗jφj⟩=−λjbjb∗j>0. |
We can see that Fd2(u,v)(dj2,0,0)Φj∉Im(L1), and so condition (c) is satisfied. The proof of (i) is completed.
(ⅱ) Suppose that there exists a positive integer i(≠j) such that di2=dj2≜ˆd. Then
kerL1(ˆd)=span{Φi,Φj},kerL∗1(ˆd)=span{Φ∗i,Φ∗j} |
and
Im(L1(ˆd))={(u,v)∈Y:∫π0(u+b∗iv)φidx=∫π0(u+b∗jv)φjdx=0}, |
which leads to codimIm(L1(ˆd))=dimkerL1(ˆd)=2.
Clearly, the Crandrall-Rabinowitz bifurcation theorem does not work in the situation since condition (b) in (ⅰ) is not satisfied. Now, we deal with this situation by the techniques of space decomposition and implicit function theorem.
To achieve our aim, we first make the following decomposition
X=X1⊕X2, |
where X1=span{Φi,Φj} and X2 is defined in (3.5). We next look for the solution of F(d2,u,v) in the following form
(u,v)=s(cosωΦi+sinωΦj+W),W=(W1,W2)T∈X2, |
where s,ω∈R are parameters. Define an operator P on Y by
P(uv)=11+bib∗i∫π0(u+b∗iv)φidxΦi+11+bjb∗j∫π0(u+b∗jv)φjdxΦj. |
Then Im(P)=span{Φi,Φj}=X1⊂Y,P2=P. Hence, P is the projection from Y to X1⊂Y. Thus we decompose Y as Y=Y1⊕Y2 with Y1:=Im(P) and Y2:=kerP=Im(L1(ˆd)).
Next, we rewrite the map F:(0,∞)×X→Y by
F(d2,u,v)=L1(d2)(uv)+(h1h2), |
where
h1(u,v)=(kv∗(mv∗+a)(a+u∗+mv∗)3−1)u2−k(amv∗+2mu∗v∗+a2+au∗)(a+u∗+mv∗)3uv+ku∗m(a+u∗)(a+u∗+mv∗)3v2+O(∣u∣3,∣u∣2∣v∣),h2(u,v)=−δ2βu∗u2+2δu∗uv−βu∗v2+O(∣u∣3,∣u∣2∣v∣). |
It is obvious that F(ˆd,0,0)=0, and F(u,v)(ˆd,0,0)=L1(ˆd). In order to verify the existence of nonconstant positive solutions of (3.2), we only need to find the existence of nonconstant pair (u,v) satisfying F(d2,u,v)=0.
Fixing ω0∈R for the time being, we define a nonlinear mapping
K(d2,s,W;ω):R+×R×X2×(ω0−δ,ω0+δ)→Y |
by
K(d2,s,W;ω)=s−1F(d2,s(cosωΦi+sinωΦj+W))=L1(d2)(cosωΦi+sinωΦj+W)+s−1(h1h2)=L1(d2)(cosωΦi+sinωΦj+W)+s(˜h1˜h2), |
where
˜h1=(−1+kv∗(mv∗+a)(a+u∗+mv∗)3)(cosωφi+sinωφj+W1)2+ku∗m(a+u∗)(a+u∗+mv∗)3(bicosωφi+bjsinωφj+W2)2−k(amv∗+2mu∗v∗+a2+au∗)(a+u∗+mv∗)3(cosωφi+sinωφj+W1)(bicosωφi+bjsinωφj+W2)+O(∣s∣), |
˜h2=−δ2βu∗(cosωφi+sinωφj+W1)2−βu∗(bicosωφi+bjsinωφj+W2)2+2δu∗(cosωφi+sinωφj+W1)(bicosωφi+bjsinωφj+W2)+O(∣s∣). |
It is clear that K(ˆd,0,0;ω0)=0. By some calculations, we see that the Frechet derivative of K(d2,s,W;ω) with respect to (d2,s,W) at (d2,s,W;ω)=(ˆd,0,0;ω0) is the liner mapping
K(d2,s,W)(ˆd,0,0;ω0)(d2,s,W)=L1(ˆd)W−d2bicosω0λi(0φi)−d2bjsinω0λj(0φj)+scos2ω0(A1φ2iB1φ2i)+scosω0sinω0(A2φiφjB2φiφj)+ssin2ω0(A3φ2jB3φ2j), |
where Ak and Bk (k=1,2,3) are given in (3.8).
We then show that
K(d2,s,W)(ˆd,0,0;ω0):R+×R×X2→Y |
is an isomorphism. To this end, we rewrite
K(d2,s,W)(ˆd,0,0;ω0)(d2,s,W)=Υ1+Υ2, |
where Υ1∈Y1,Υ2∈Y2, and we decompose
(0φi)=c1Φi+(u1v1)and(0φj)=c2Φj+(u2v2), |
where
c1=b∗i1+bib∗i≠0,(u1v1)=(−c11−c1bi)φi; |
c2=b∗j1+bjb∗j≠0,(u2v2)=(−c21−c2bj)φj. |
Furthermore, we can easily check that
(u1v1),(u2v2)∈Y2. |
In the next moment, we shall divide our discussion into two cases j=2i and i=2j.
Case 1: j=2i. In this case, a simple calculation yields
∫π0φ2iφj=√12π,∫π0φ2jφi=0and∫π0φ3i=∫π0φ3j=0. |
Then, it is clear that
(A3φ2jB3φ2j)∈Y2. |
We decompose
(A1φ2iB1φ2i)=c3Φj+(u3v3)and(A2φiφjB2φiφj)=c4Φi+(u3v3), |
where
c3=A1+B1b∗j1+bjb∗j∫π0φ2iφjdx=√12πA1+B1b∗j1+bjb∗j,(u3v3)=(A1φ2i−c3φjB1φ2i−c3bjφj)∈Y2, |
c4=A2+B2b∗i1+bib∗i∫π0φ2iφjdx=√12πA2+B2b∗i1+bib∗i,(u4v4)=(A2φiφj−c4φiB2φiφj−c4biφi)∈Y2. |
By the decomposition of Y, we have
Υ1=(−d2c1biλicosω0+sc4sinω0cosω0)Φi+(−d2c2bjλjsinω0+sc3cos2ω0)Φj,Υ2=L1(ˆd)W−d2biλicosω0(u1v1)−d2bjλjsinω0(u2v2)+scos2ω0(u3v3)+scosω0sinω0(u4v4)+ssin2ω0(A3φ2jB3φ2j). |
Let K(d2,s,W)(ˆd,0,0;ω0)(d2,s,W)=0, then we get Υ1=0 and Υ2=0. Since ω0 satisfies (3.6) when j=2i, we obtain d2=0 and s=0 from Υ1=0. Embedding them into Υ2=0, we have W=0. This shows that K(d2,s,W)(ˆd,0,0;ω0) is injective.
We now prove that K(d2,s,W)(ˆd,0,0;ω0) is surjective. For any (u,v)∈Y, we need to find (d2,s,W)∈R+×R×X2 such that
K(d2,s,W)(ˆd,0,0;ω0)(d2,s,W)=(uv). | (3.10) |
By the decomposition of Y, there exist α,β∈R and (ˉu,ˉv)∈Y2 such that
(uv)=αΦi+βΦj+(ˉuˉv). |
Substituting it into (3.10), we obtain
{−d2c1biλicosω0+sc4sinω0cosω0=α,−d2c2bjλjsinω0+sc3cos2ω0=β,L1(ˆd)W−d2biλicosω0(u1v1)−d2bjλjsinω0(u2v2)+scos2ω0(u3v3)+scosω0sinω0(u4v4)+ssin2ω0(A3φ2jB3φ2j)=(ˉuˉv). | (3.11) |
By (3.6), we obtain
d2=~d2:=βc4sinω0−αc3cosω0c1c3biλicos2ω0−c2c4bjλjsin2ω0,s=˜s:=αc2bjλjsinω0−βc1biλicosω0c2c4bjλjsin2ω0cosω0−c1c3biλicos3ω0. |
Note that L1(ˆd) is an isomorphism from X2 to Y2, and we get
W=L−11(ˆd)(˜u˜v), |
by embedding d2=~d2 and s=˜s into the third equation of (3.11), where
(˜u˜v)=(ˉuˉv)+~d2biλicosω0(u1v1)+~d2bjλjsinω0(u2v2)−˜scos2ω0(u3v3) |
−˜scosω0sinω0(u4v4)−˜ssin2ω0(A3φ2jB3φ2j). |
Then
(d2,s,W)=(~d2,˜s,L−11(ˆd)(˜u˜v)) |
is the solution of (3.10), which implies K(d2,s,W)(ˆd,0,0;ω0) is surjective.
Applying the implicit function theorem to
K(d2,s,W;ω)=0, | (3.12) |
we can know that there is a curve of nonconstant solutions (d2(ω),s(ω),W(ω)) of (3.12) in small neighborhood of ω0, where ω0 satisfies (3.6), d2(ω),s(ω) and W(ω) are continuously differentiable functions and satisfy d2(ω0)=ˆd,s(ω0)=0,W(ω0)=0. Therefore, (d2(ω),s(ω)(cosωΦi+sinωΦj)+W(ω)) are nonconstant solutions of F(d2,(u,v))=0.
Case 2: i=2j. A simple calculation yields
∫π0φ2iφj=0,∫π0φ2jφi=√12πand∫π0φ3i=∫π0φ3j=0. |
Then
(A1φ2iB1φ2i)∈Y2. |
We decompose
(A3φ2jB3φ2j)=c5Φi+(u5v5)and(A2φiφjB2φiφj)=c6Φj+(u6v6), |
where
c5=A3+B3b∗i1+bib∗i∫π0φiφ2jdx=√12πA3+B3b∗i1+bib∗i,(u5v5)=(A3φ2j−c5φiB3φ2j−c5biφi)∈Y2; |
c6=A2+B2b∗j1+bjb∗j∫π0φiφ2jdx=√12πA2+B2b∗j1+bjb∗j,(u6v6)=(A2φiφj−c6φjB2φiφj−c6bjφj)∈Y2. |
Hence, we have
Υ1=(−d2c1biλicosω0+sc5sin2ω0)Φi+(−d2c2bjλjsinω0+sc6sinω0cosω0)Φj,Υ2=L1(ˆd)W−d2biλicosω0(u1v1)−d2bjλjsinω0(u2v2)+scos2ω0(A1φ2iB1φ2i)+scosω0sinω0(u6v6)+ssin2ω0(u5v5). |
As in Case 1 above, if ω0 satisfies (3.7), then K(d2,s,W)(ˆd,0,0;ω0) is an isomorphism from :R+×R×X2 to Y. By the implicit function theorem, we finish the proof of this case. Thus the whole proof is completed.
Remark 3.2. In Theorem 3.1 (ii), the existence of nonconstant positive solution of (3.2) is showed. Indeed, we derive many curves of nonconstant solutions because ω0 is not unique and can be chosen only to satisfy (3.6) or (3.7). Moreover, due to bi≠bj, it is impossible that both Ak and Bk,k=1,2,3, are simultaneously equal to zero, which implies that there must exists ω0 satisfying the condition (3.6) above. Similarly, we can check that there must exists ω0 satisfying the condition (3.7) above.
Theorem 3.1 provides no information of the bifurcation curve Γj far from the equilibrium. In order to understand its global structure, a further study is therefore necessary.
We first introduce the standard abstract bifurcation theorem from [37] for readers' convenience. Let X be a Banach space and let T:R×X→X be a compact, continuously differentiable operator such that T(a,0)=0. Assume that T can be written as
T(a,U)=K(a)U+W(a,U), | (3.13) |
where K(a) is a linear compact operator and the Fréchet derivative WU(a,0)=0. Regarding a as a bifurcation parameter, we will undertake a global bifurcation analysis for the equation
U=T(a,U). | (3.14) |
We suppose that I−K:X→X is a bijection. Then the Leray-Schauder degree
deg(I−K,ˆB,0)=(−1)p, |
where ˆB is a ball centered at 0 in X and p is the sum of the algebraic multiplicities of the eigenvalues of K that are larger than 1. If x0 is an isolated fixed point of the operator T and B is a ball centered at x0 such that x0 is the unique fixed point of T in B, the index of T at x0 is defined as
index(T,x0)=deg(I−T,B,x0). |
Moreover, if x0 is a fixed point of T and I−T∗(x0) is invertible, then x0 is an isolated fixed point of T and
index(T,x0)=deg(I−T,B,x0)=deg(I−T∗(x0),ˆB,0), |
where B and ˆB are sufficiently small.
We now state the result on the global bifurcation for the operator T defined by (3.14).
Lemma 3.3. [37, Theorem 1.3] Let a0 be such that I−K(a) is invertible if 0<|a−a0|<ϵ for ϵ>0. Assume that index(T(a,⋅),0) is constant on (a0−ϵ,a0) and on (a0,a0+ϵ); moreover, if a0−ϵ<a1<a0<a2<a0+ϵ, then index(T(a1,⋅),0)=index(T(a2,⋅),0). Then there exists a continuum C in the a−U plane of solutions of (3.14) such that one of the following alternatives is true
(i) C joins (a0,0) to (ˆa,0) where I−K(ˆa) is not invertible;
(ii) C joins (a0,0) to ∞ in R×X.
Theorem 3.4. Under the same assumptions as in Theorem 3.1, the projection of the bifurcation curve Γj is on the d2-axis contains (dj2,∞). If d2>¯d2 and d2≠dk2 for any integer k>0, then model (1.2) possesses at least one non-constant positive solution, where ¯d2=min0≤i≤i0di2.
Proof. First, we rewrite model (1.2) in a form that the standard global bifurcation theory can be more conveniently applied.
Let ˜u=u−u∗,˜v=v−v∗, then (1.2) is transformed into
{−d1˜uxx=a11˜u+a12˜v+h1(˜u,˜v),−d2˜vxx=a21˜u+a22˜v+h2(˜u,˜v), | (3.15) |
where h1(˜u,˜v),h2(˜u,˜v) are higher-order terms of ˜u and ˜v. The constant steady state (u∗,v∗) of (1.2) shifts to (0,0) of this new system.
Let G1:h→ω denote the Green operator for the boundary value problem
a11ω−d1ωxx=hin(0,π),ωx=0on0,π, |
and G2:h→ω the Green operator for
a22ω−d2ωxx=hin(0,π),ωx=0on0,π, |
where a11>0 and a22<0. Put ˜U=(˜u,˜v),
K(d2)˜U=(2a11G1(˜u)+a12G1(˜v),a21G2(˜u)) |
and
H(˜U)=(G1(h1(˜u,˜v)),G2(h2(˜u,˜v)). |
Recall that
X={(u,v)∣u,v∈C2(0,π),ux=vx=0,x=0,π}. |
Then the boundary value problem (3.15) can be interpreted as the equation
˜U=K(d2)˜U+H(˜U)inX. | (3.16) |
Note that K(d2) is a compact liner operator on X for any given d2>0,H(˜U)=o(|˜U|) for ˜U near zero uniformly on closed d2 sub-intervals of (0,∞), and is a compact operator on X as well.
In order to apply Rabinowitz's global bifurcation theorem, we first verify that 1 is an eigenvalue of K(dj2) of algebraic multiplicity one. From the argument in the proof of Theorem 3.1 it is seen that ker(K(dj2)−I))=kerL1=span{Φj}, so 1 is indeed an eigenvalue of K=K(dj2), and dimker(K−I)=1. As the algebraic multiplicity of the eigenvalue 1 is the dimension of the generalized hull space ∪∞i=1ker(K−I)i, we need to verify that ker(K−I)=ker(K−I)2, or ker(K−I)∩R(K−I)=0.
We now compute ker(K∗−I) following the calculation in [36], where K∗ is the adjoint of K.
Let (ˆϕ,ˆψ)∈ker(K∗−I), then
2a11G1(ˆϕ)+a22G2(ˆψ)=ˆϕ,a12G1(ˆϕ)=ˆψ. |
By the definition of G1 and G2 we obtain
−dj2a12∂2∂x2ˆϕ=fˆϕˆϕ+fˆψˆψ,−d1∂2∂x2ˆψ=a12ˆϕ−a11ˆψ, |
where
fˆϕ=2dj2a11a22d1+a12a21,fˆψ=a12a21−2(a11a22+dj2a211d1). |
Let ˆϕ=∑∞i=0ˆaiφi,ˆψ=∑∞i=0ˆbiφi, then
∞∑i=0ˆBi(ˆaiˆbi)φi=0,whereˆBi=(−dj2a12λi+fˆϕfˆψa12−d1λi−a11). |
By a straightforward calculation one can check that detˆBi=a12detˉBi, where ˉBi is given in (3.9) (replacing d2 there by dj2). Thus detˉBi=0 only for i=j, and ker(K∗−I)=span{ˆΦj}, where ˆΦj=(d1λi+a11,1)⊤φj. This shows that Φj∉(ker(K∗−I))⊥=R(K−I), so ker(K−I)∩R(K−I)=0 and the eigenvalue 1 has algebraic multiplicity one.
If 0<d2≠dj2 is in a small neighborhood of dj2, then the liner operator I−K(d2):X→X is a bijection and 0 is an isolated solution of (3.15) for this fixed d2. The index of this isolated zero of I−K(d2)−H is given by
index(I−K(d2)−H,(d2,0))=deg(I−K(d2),B,0)=(−1)p, |
where B is a sufficiently small ball center at 0, and p is the sum of the algebraic multiplicities of the eigenvalues of K(d2) lager than 1.
For our bifurcation analysis, it is also necessary to show that this index changes as d2 crosses dj2, that is, for ϵ>0 sufficiently small, we need to verify
index(I−K(dj2−ϵ)−H,(dj2−ϵ,0))≠index(I−K(dj2+ϵ)−H,(dj2+ϵ,0)). | (3.17) |
Indeed, if μ is an eigenvalue of K(d2) with an eigenfunction (ˆϕ,ˆψ), then
−d1μˆϕxx=(2−μ)a11ˆϕ+a12ˆψ,−d2μˆψxx=a21ˆϕ+a22μˆψ. |
By using the Fourier cosine series ˆϕ=Σˆaiφi and ˆψ=Σˆbiφi, we have
∞∑i=0~Bi(ˆaiˆbi)φi=0,where~Bi=((2−μ)a11−d1λiμa12a21(a22−d2λi)μ). |
Thus the set of eigenvalues of K(d2) consists of all μ which implies the following characteristic equation
(d1λi+a11)μ2−2a11μ−a12a21d2λi−a22=0, | (3.18) |
where the integer i runs from zero to ∞. In particular, for d2=dj2, if μ=1 is a root of (3.18), then a simple calculation leads to dj2=di2, and so j=i by the assumption. Therefore, without counting the eigenvalues corresponding to i≠j in (3.18), K(d2) has the same number of eigenvalues greater than 1 for all d2 close to dj2, and they have the same multiplicities. On the other hand, for i=j in (3.18), we let μ(d2),˜μ(d2) denote the two roots of (3.18). By a straightforward calculation, we find that
μ(dj2)=1and˜μ(dj2)=a11−d1λia11+d1λi<1. |
Now for d2 close to dj2, we obtain ˜μ(dj2)<1. As the constant term −a12a21/d2λi−a22 in (3.18) is a decreasing function of d2, there exists
μ(dj2+ϵ)>1,μ(dj2−ϵ)<1. |
Consequently, K(dj2+ϵ) has exactly one more eigenvalues that are larger than 1 than K(dj2−ϵ) does, and by a similar argument above we can show this eigenvalue has algebraic multiplicity one. This verifies (3.17).
Therefore, we apply Theorem 1.3 in [37] to conclude that Γj either
(1) meets infinity in R×X, or
(2) meets (dk2,(u∗,v∗)) for some k≠j,dk2>0.
We show that Γj must extend to infinity in R×X by the idea of Nishiura [38] and Takagi [39]. Thus, the theorem is verified.
Remark 3.5. Theorem 3.4 shows that there is a smooth curve Γj of positive solutions of model (1.2) bifurcating from (dj2,u∗,v∗), with Γj contained in a global branch of the positive solutions of (1.2).
In this paper, based on the results in [17], we make a detailed descriptions for the local (c.f., Theorem 3.1) and global (c.f., Theorem 3.4) bifurcation structure of nonconstant positive steady states of a modified Holling-Tanner predator-prey system under homogeneous Neumann boundary condition. These theorems and the results in [17] can give a profile of the solutions of model (1.1). The results are beneficial to population persistence control, that is, we must do our best to regulate the parameters in the special range to avoid population extinction.
It is should be noted that our results are based on 1-dimensional space. And in the N-dimensional space, N≥2, we can also obtain a local bifurcation result, which is an analogue of Theorem 3.1. But the global bifurcation (Theorem 3.4) is only established in the square domain. For instance, in the special 2-D case with (0,L)×(0,L). In this special case, the Neumann eigenvalue problem has eigen-pairs
λmn=(m2+n2)π2L2,φmn(x,y)=cosmπxLcosnπyL,m,n∈N+. |
Each eigenvalue gives rise to bifurcation point dmn2, with λj being replaced by λmn in (3.3). We mention that though the boundary of the square is not smooth, the Neumann boundary condition has to be interpreted in the weak fashion via first Green's identity in the standard way [40].
This research was supported by the National Science Foundation of China (11601179, 61672013 and 61772017) and Huaian Key Laboratory for Infectious Diseases Control and Prevention (HAP201704), Huaian, Jiangsu Province, China.
The authors declare that they have no competing interests.
[1] | J. Jang, W.-M. Ni and M. Tang, Global bifurcation and structure of Turing patterns in the 1-D Lengyel-Epstein model, J. Dyn. Diff. Equa., 2 (2004), 297–320. |
[2] | A. M. Turing, The chemical basis of morphogenesis, Philos. T. R. SOC. B, 237 (1852), 37–72. |
[3] | N. F. Britton, Essential Mathematical Biology, Springer, New York, 2003. |
[4] | L. A. Segel and J. L. Jackson, Dissipative structure: an explanation and an ecological example, J. Theor. Biol., 37 (1972), 545–559. |
[5] | D. Alonso, F. Bartumeus and J. Catalan, Mutual interference between predators can give rise to Turing spatial patterns, Ecology, 83 (2002), 28–34. |
[6] | W.-M. Wang, L. Zhang, H. Wang, et al., Pattern formation of a predator-prey system with Ivlev-type functional response, Ecol. Model., 221 (2008), 131–140. |
[7] | C. Neuhauser, Mathematical challenges in spatial ecology, Notices AMS, 48 (2001), 1304–1314. |
[8] | A. Okubo and S. A. Levin, Diffusion and Ecological Problems: Modern Perspectives, Springer, New York, 2001. |
[9] | A. B. Medvinsky, S. V. Petrovskii, I. A. Tikhonova, et al., Spatiotemporal complexity of plankton and fish dynamics. SIAM Rev., 44 (2002), 311–370. |
[10] | J. D. Murray, Mathematical biology. II: Spatial models and biomedical applications, Springer, New York, 2003. |
[11] | P. Y. H. Pang and M. Wang, Qualitative analysis of a ratio-dependent predator-prey system with diffusion, Proc. R. Soc. Edinb., 133 (2003), 919–942. |
[12] | K. Kuto and Y. Yamada. Multiple coexistence states for a prey-predator system with crossdi ffusion, J. Differ. Equations, 197 (2004), 315–348. |
[13] | K. Kuto, Stability of steady-state solutions to a prey-predator system with cross-diffusion, J. Differ. Equations, 197 (2004), 293–314. |
[14] | X. Zeng and Z. Liu, Non-constant positive steady states of a prey-predator system with cross-diffusions, J. Math. Anal. Appl., 332 (2007), 989–1009. |
[15] | R. Peng and J. Shi, Non-existence of non-constant positive steady states of two Holling type-II predator-prey systems: Strong interaction case, J. Differ. Equations, 247 (2009), 866–886. |
[16] | F. Yi, J. Wei and J. Shi, Bifurcation and spatiotemporal patterns in a homogeneous diffusive predator-prey system, J. Differ. Equations, 246 (2009), 1944–1977. |
[17] | H. Shi,W.-T. Li and G. Lin. Positive steady states of a diffusive predator-prey system with modified Holling-Tanner functional response, Nonl. Anal. Real, 11 (2010), 3711–3721. |
[18] | Y. Cai, M. Banerjee, Y. Kang, et al., Spatiotemporal complexity in a predator-prey model with weak Allee effects, Math. Biosci. Eng., 11 (2014), 1247–1274. |
[19] | S. Li, J. Wu and Y. Dong, Turing patterns in a reaction-diffusion model with the Degn-Harrison reaction scheme, J. Differ. Equations, 259 (2015), 1990–2029. |
[20] | H. Shi and S. Ruan. Spatial, temporal and spatiotemporal patterns of diffusive predator-prey models with mutual interference, IMA J. Appl. Math., 80 (2015), 1534–1568. |
[21] | Y. Cai and W.-M. Wang, Fish-hook bifurcation branch in a spatial heterogeneous epidemic model with cross-diffusion, Nonl. Anal. Real, 30 (2016), 99–125. |
[22] | T. Kuniya and J.Wang, Global dynamics of an SIR epidemic model with nonlocal diffusion, Nonl. Anal. Real, 43 (2018), 262–282. |
[23] | J.Wang, J.Wang and T. Kuniya, Analysis of an age-structured multi-group heroin epidemic model.Appl. Math. Comp., 347 (2019), 78–100. |
[24] | Y. Cai, Z. Ding, B. Yang, et al., Transmission dynamics of Zika virus with spatial structure–A case study in Rio de Janeiro, Brazil. Phys. A, 514 (2019), 729–740. |
[25] | Y. Cai, K. Wang and W.M. Wang, Global transmission dynamics of a Zika virus model, Appl. Math. Lett., 92 (2019), 190–195. |
[26] | Y. Cai, X. Lian, Z. Peng, et al., Spatiotemporal transmission dynamics for influenza disease in a heterogenous environment, Nonl. Anal. Real., 46 (2019), 178–194. |
[27] | Y. Cai, Z. Gui, X. Zhang, et al., Bifurcations and pattern formation in a predator-prey model,. Inter. J. Bifur. Chaos, 28 (2018), 1850140. |
[28] | H. Zhang, Y. Cai, S. Fu, et al., Impact of the fear effect in a prey-predator model incorporating a prey refuge, Appl. Math. Comp., 356 (2019), 328–337. |
[29] | X. Cao, Y. Song and T. Zhang. Hopf bifurcation and delay-induced Turing instability in a diffusive Iac Operon model, Inter. J. Bifur. Chaos, 26 (2016), 1650167. |
[30] | J. Jiang, Y. Song and P. Yu, Delay-induced Triple-Zero bifurcation in a delayed Leslie-type predator-prey model with additive Allee effect, Inter. J. Bifur. Chaos, 26 (2016), 1650117. |
[31] | Y. Song, H. Jiang, Q. Liu, et al., Spatiotemporal dynamics of the diffusive Mussel-Algae model near Turing-Hopf bifurcation. SIAM J. Appl. Dyn. Sys., 16 (2017), 2030–2062. |
[32] | Y. Song and X. Tang, Stability, steady-state bifurcations, and Turing patterns in a predator-prey model with herd behavior and prey-taxis, Stud. Appl. Math., 139 (2017), 371–404. |
[33] | S. Wu and Y. Song, Stability and spatiotemporal dynamics in a diffusive predator Cprey model with nonlocal prey competition, Nonl. Anal. Real, 48 (2019), 12–39. |
[34] | S.-B. Hsu and T.-W. Huang, Global stability for a class of predator-prey systems, SIAM J. Appl. Math., 55 (1995), 763–783. |
[35] | Y. Lou and W.-M. Ni. Diffusion, self-diffusion and cross-diffusion, J. Differ. Equations, 131 (1996), 79–131. |
[36] | W.-M. Ni and M. Tang, Turing patterns in the Lengyel-Epstein system for the CIMA reaction, T. Am. Math. Soc., 357 (2005), 3953–3969. |
[37] | P. H. Rabinowitz, Some global results for nonlinear eigenvalue problems, J. Func. Anal., 7 (1971), 487–513. |
[38] | Y. Nishiura, Global structure of bifurcating solutions of some reaction-diffusion systems, SIAM J. Math. Anal., 13 (1982), 555–593. |
[39] | I. Takagi, Point-condensation for a reaction-diffusion system, J. Differ. Equations, 61 (1986), 208–249. |
[40] | A.Chertock, A. Kurganov, X. Wang, et al., On a chemotaxis model with saturated chemotactic flux, Kinet. Relat. Models, 5 (2012), 51–95. |
1. | Xiaoni Wang, Gaihui Guo, Jian Li, Mengmeng Du, Qifeng Zhang, Qualitative Analysis of a Three-Species Reaction-Diffusion Model with Modified Leslie-Gower Scheme, 2021, 2021, 2314-8888, 1, 10.1155/2021/6650783 | |
2. | Debasis Mukherjee, Dynamics of a diffusive two predators- one prey system , 2023, 1, 2980-2474, 47, 10.61383/ejam.20231349 |