In this paper, we investigate a predator-prey system with fractional type cross-diffusion incorporating the Beddington-DeAngelis functional response subjected to the homogeneous Neumann boundary condition. First, by using the maximum principle and the Harnack inequality, we establish a priori estimate for the positive stationary solution. Second, we study the non-existence of non-constant positive solutions mainly by employing the energy integral method and the Poincaré inequality. Finally, we discuss the existence of non-constant positive steady states for suitable large self-diffusion d2 or cross-diffusion d4 by using the Leray-Schauder degree theory, and the results reveal that the diffusion d2 and the fractional type cross-diffusion d4 can create spatial patterns.
Citation: Pan Xue, Cuiping Ren. Spatial patterns for a predator-prey system with Beddington-DeAngelis functional response and fractional cross-diffusion[J]. AIMS Mathematics, 2023, 8(8): 19413-19426. doi: 10.3934/math.2023990
[1] | Hong Qiu, Yanzhang Huo, Tianhui Ma . Dynamical analysis of a stochastic hybrid predator-prey model with Beddington-DeAngelis functional response and Lévy jumps. AIMS Mathematics, 2022, 7(8): 14492-14512. doi: 10.3934/math.2022799 |
[2] | Jorge Luis Ramos-Castellano, Miguel Angel Dela-Rosa, Iván Loreto-Hernández . Bifurcation analysis for the coexistence in a Gause-type four-species food web model with general functional responses. AIMS Mathematics, 2024, 9(11): 30263-30297. doi: 10.3934/math.20241461 |
[3] | Francesca Acotto, Ezio Venturino . How do predator interference, prey herding and their possible retaliation affect prey-predator coexistence?. AIMS Mathematics, 2024, 9(7): 17122-17145. doi: 10.3934/math.2024831 |
[4] | Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045 |
[5] | Yaping Wang, Yuanfu Shao, Chuanfu Chai . Dynamics of a predator-prey model with fear effects and gestation delays. AIMS Mathematics, 2023, 8(3): 7535-7559. doi: 10.3934/math.2023378 |
[6] | Changtong Li, Dandan Cheng, Xiaozhou Feng, Mengyan Liu . Complex dynamics of a nonlinear impulsive control predator-prey model with Leslie-Gower and B-D functional response. AIMS Mathematics, 2024, 9(6): 14454-14472. doi: 10.3934/math.2024702 |
[7] | Shanshan Yu, Jiang Liu, Xiaojie Lin . Multiple positive periodic solutions of a Gause-type predator-prey model with Allee effect and functional responses. AIMS Mathematics, 2020, 5(6): 6135-6148. doi: 10.3934/math.2020394 |
[8] | Jagdev Singh, Behzad Ghanbari, Ved Prakash Dubey, Devendra Kumar, Kottakkaran Sooppy Nisar . Fractional dynamics and computational analysis of food chain model with disease in intermediate predator. AIMS Mathematics, 2024, 9(7): 17089-17121. doi: 10.3934/math.2024830 |
[9] | Kimun Ryu, Wonlyul Ko . Stability and bifurcations in a delayed predator-prey system with prey-taxis and hunting cooperation functional response. AIMS Mathematics, 2025, 10(6): 12808-12840. doi: 10.3934/math.2025576 |
[10] | Chuangliang Qin, Jinji Du, Yuanxian Hui . Dynamical behavior of a stochastic predator-prey model with Holling-type III functional response and infectious predator. AIMS Mathematics, 2022, 7(5): 7403-7418. doi: 10.3934/math.2022413 |
In this paper, we investigate a predator-prey system with fractional type cross-diffusion incorporating the Beddington-DeAngelis functional response subjected to the homogeneous Neumann boundary condition. First, by using the maximum principle and the Harnack inequality, we establish a priori estimate for the positive stationary solution. Second, we study the non-existence of non-constant positive solutions mainly by employing the energy integral method and the Poincaré inequality. Finally, we discuss the existence of non-constant positive steady states for suitable large self-diffusion d2 or cross-diffusion d4 by using the Leray-Schauder degree theory, and the results reveal that the diffusion d2 and the fractional type cross-diffusion d4 can create spatial patterns.
In this paper, we investigate the following predator-prey system with cross-diffusion incorporating the Beddington-DeAngelis functional response subjected to the homogeneous Neumann boundary condition
{ ut−d1Δ[(1+d3v)u]=u(1−u−v1+au+bv),(x,t)∈Ω×(0,∞),vt−d2Δ[(1+d41+u)v]=v(−d+cu1+au+bv),(x,t)∈Ω×(0,∞),u(x,0)=u0(x)≥0,v(x,0)=v0(x)≥0,x∈Ω,∂νu=∂νv=0,(x,t)∈∂Ω×(0,∞), | (1.1) |
where Ω⊂RN is a bounded domain with smooth boundary ∂Ω, N≥1 is an integer, Δ=∑Ni=1 is the Laplace operator in RN, ν represents the outward unit normal vector on the boundary ∂Ω with ∂ν=∂∂ν, and the homogeneous Neumann boundary condition means that the individuals do not cross the habitat boundary, u and v represent the densities of prey and predator, respectively. The parameters a,b,c and d are all positive constants. The interaction between the prey and the predator of system (1.1) is the most usually used the Beddington-DeAngelis functional response
p(u,v)=u1+au+bv, |
which was introduced by Beddington and DeAngelis, where, the parameters a,b,c>0 are the saturation constant for an alternative prey, the predator interference and the consumption rate, respectively. The term au in the denominator describing mutual interference among the preys while the term bv describing that among the predators. It is well known that the Beddington-DeAngelis functional response has desirable qualitative features of ratio-dependent form but takes care of their controversial behaviors at low densities. Compared with Holling-II functional response, Beddington and DeAngelis response, which considered both the mutual interference among the predator and the handing time of each prey, is more reasonable. One can refer to [1,2,3,4] for more details on the background of this functional response.
In view of the inhomogeneous distribution of the predator and prey in different spatial locations with a fixed domain Ω at any given time and the natural tendency of each species to diffuse, we take into account the predator-prey system (1.1), with self- and cross-diffusions. The role of diffusion into the modelling has been extensively studied. Generally speaking, the diffusion process usually gives rise to a stabilizing effect so that generates a constant equibrium state, namely, the spatial pattern of morphogen or chemical concentration. System (1.1) implies that, in addition to the dispersive force, the diffusion also depends population pressure from other species. The flux of diffusion to the predators of the system is
−▽(1+d41+u)v=−(1+d41+u)▽v+d4v(1+u)2▽u. |
The part d4v(1+u)2▽u of the diffusion flux is directed toward the increasing densities of the prey, which implies the preys respond to attack of team for the movement of predators. The part −(1+d41+u)▽v of the diffusion flux is directed toward the decreasing densities of the predators, which indicates that the predators move towards the preys. The interplay between these diffusion terms and the population dynamics given by (1.1) can lead to complex spatial patterns in the predator-prey system. For example, it may be possible for regions with high predator density to drive local extinction of the prey, leading to further reduction in predator density and eventual recolonization of the area by prey. This kind of spatial dynamics is often observed in natural ecological systems, and can have important implications for conservation efforts and management of ecosystems. For more details about the biological significance, one can see [5,6,7,14,15] for references, in which the predator-prey system with self- and cross-diffusions were considered.
In [8], by employing the Fixed point index theory, the authors studied the existence of the non-constant steady state of a predator-prey with the Beddington-DeAngelis functional response, in which the cross-diffusions are linear. Paper [9] considers the stationary problem of the Holling-Tanner prey-predator model with fractional type cross-diffusion terms, and result reveals that the large cross-diffusion can create spatial patterns. Papers [10,11] mainly consider the existence of the non-constant positive solutions by making use of the Leray-Schauder degree theory, Furthermore, the authors also discussed the Turing instability of a Gause-type predator-prey system with self-and cross-diffusions mainly by considering the influence of the diffusion terms.
Although lots of researchers have investigated the predator-prey system with nonlinear diffusions(see[12,13,14,15] for reference), there are still many open problem on the spatial patterns cased by the linear and nonlinear cross-diffusions. To our knowledge, there are few works focused on system (1.1), which included the Beddington-DeAngelis functional response and a fractional cross-diffusion for the predator. Paper [16] considers a strongly coupled partial differential equation model with a non-monotonic functional response—a Holling type-IV function in a bounded domain with no flux boundary condition. The authors proved a number of existence and non-existence results concerning non-constant steady states of the underlying system. The main purpose of this paper is to research into the effect of the self- and cross-diffusions on the non-constant positive solutions of system (1.1), namely, we investigate the existence and non-existence of the non-constant positive solutions to the following elliptic system
{ −d1Δ[(1+d3v)u]=u(1−u−v1+au+bv),x∈Ω,−d2Δ[(1+d41+u)v]=v(−d+cu1+au+bv),x∈Ω,∂νu=∂νv=0,x∈∂Ω. | (1.2) |
In system (1.2), the cross-diffusions implies that the movement of the species at any spatial location is influenced by the gradient of the concentration of the interacting species at that location. By taking these facts into account, the system can capture much more richer phenomena, and this deserves our careful study and discussion.
The organization of this article is as follows. In Section 2, we give a priori estimate for the positive stationary solution by using the maximum principle and the Harnack inequality. In Section 3, we study the non-existence of non-constant positive solutions mainly by employing the energy integral method and the Poincaré inequality. Moreover, we also discuss the existence of non-constant positive steady states for suitable self- and cross-diffusion coefficients by employing the Leray-Schauder degree theory. The results reveal that the diffusion d2 and the fractional type cross-diffusion d4 can create spatial patterns.
We know that there exist three non-negative constant solutions (0,0), (1,0) and (u∗,v∗) for system (1.1), where
u∗=12bc[ad+bc−c+√(ad+bc−c)2+4bcd], v∗=cd(1−u∗)u∗, |
and u∗<1 provided by the condition c>d(1+a). Therefore, it is necessary to assume that c>d(1+a) holds throughout this paper so as to (u∗,v∗) is the unique positive constant solution of system (1.2). For convenience, we denote Θ=Θ(a,b,c,d) in the sequel.
In this section, in order to obtain a priori estimates of the positive solution of system (1.2), we first present the following lemmas, named the Maximum principle and Harnack inequality[17,18], respectively.
Lemma 2.1. Let ϕ(u,x)∈C(Ω×R1). If u∈C2(Ω)∩C1(¯Ω) satisfies Δu(x)+ϕ(u,x)≥0 in Ω, ∂νu=0 on ∂Ω and achieves its maximum at a point x0∈¯Ω, then −Δu(x0)≥0.
Lemma 2.2. Assume that c(x)∈C(¯Ω) and let ω(x)∈C2(Ω)∩C1(¯Ω) be a positive solution to
Δω(x)+c(x)ω(x)=0,x∈Ω,∂νω=0,x∈∂Ω. |
Then there exists a positive constant C∗=C∗(||c||∞,Ω), such that max¯Ωω≤C∗min¯Ωω.
Theorem 2.3. Let C1, D1,D2,D3 be given positive constants. Then there exists a positive constant C=C(Θ,D1,D2,D3,Ω) which is independent of di(i=1,2,3,4), such that every positive solution (u,v) of system (1.2), satisfies C−1<u(x),v(x)<C for d1≥D1, d2≥D2, d3≤D3, C1>1a+c and b<1.
Proof. Let ϕ1=d1[(1+d3v)u],ϕ2=d2[(1+d41+u)v]. Assume that there exists a point x0∈¯Ω such that ϕ1(x0)=max¯Ωϕ1. According to Lemma 2.1, we obtain u(x0)<1,v(x0)≤1+a1−b, then
max¯Ωu≤1d1max¯Ωϕ1≤1d1(1+d3v(x0))u(x0)≤1D1(1+D31+a1−b)≜C1, |
where b<1, D1 and D3 are given positive constants with d1≥D1, d3≤D3. Let x1∈¯Ω be a point such that ϕ2(x1)=max¯Ωϕ2. Then, by the maximum principle, we have
v(x1)≤(a+c)C1−1b, |
where C1>1a+c. Therefore, we can obtain
max¯Ωv≤1d2max¯Ωϕ2min¯Ω(1+d41+u(x))=1+d41+d41+C1v(x1)≤(1+C1)(a+c)C1−1b≜C2. |
as 1+d41+d41+C1 is strictly increasing with respect to d4 and satisfies limd4→+∞1+d41+d41+C1=1+C1. Thus, we obtain the upper bounds of the solution (u,v).
Hereinafter, we show that (u,v) has a lower bound. For convenience, we set
c1(x)=1d1(1+d3v)(1−u−v1+au+bv), c2(x)=1d2(1+d41+u)(−d+cu1+au+bv). |
Then, system (1.2) can be written as
{−Δϕ1(x)=c1(x)ϕ1(x),x∈Ω,−Δϕ2(x)=c2(x)ϕ2(x),x∈Ω,∂νϕ1=∂νϕ2=0,x∈∂Ω. | (2.1) |
Since ||c1(x)||∞<C3(Θ,D1,D2,D3,Ω), according to Lemma 2.2, we know that there exists a positive constant C4=C4(Θ,D1,D2,D3,Ω) and C5=C5(Θ,D1,D2,D3,Ω) such that
max¯Ωϕ1≤C4min¯Ωϕ1, |
and
max¯Ωumin¯Ωu≤max¯Ωϕ1min¯Ωϕ1⋅1+d3max¯Ωv1+d3min¯Ωv≤C4(1+d3max¯Ωv)≤C4(1+D3C2)≜C5. | (2.2) |
Similarly, as ||c2(x)||∞<1D2(d+ca)≜C6, Lemma 2.2 holds for ϕ2, that is,
max¯Ωϕ2≤C7min¯Ωϕ2, |
for some positive constant C7. Therefore, we have
max¯Ωvmin¯Ωv≤max¯Ωϕ2min¯Ωϕ2⋅max¯Ω(1+d41+u)min¯Ω(1+d41+u)≤C71+d41+d41+C1≤C7(1+C1)≜C8, | (2.3) |
for some positive constant C8. Thus, if the positive solution (u,v) does not have positive lower bound, then there exists a sequence {(d1i,d2i,d3i,d4i)} satisfying d1i≥D1,d2i≥D3 and d3i≤D3 such that the corresponding solutions of system (1.2) satisfy min¯Ωui→0, or min¯Ωvi→0 as i→∞. Due to the Harnack inequality, we have max¯Ωui→0, or max¯Ωvi→0 as i→∞.
By integrating the second equation of system (1.2) over Ω, we obtain ∫Ωv(−d+cu1+au+bv)dx=0. Thus, there must exists a point x1∈¯Ω such that cu(x1)1+au(x1)+bv(x1)=d, which implies u(x1)≥dc. Thus, combined with (2.2), we have
min¯Ωu≥max¯ΩuC5=dcC5. |
Similarly, by integrating the first equation of system (1.2) over Ω, we obtain
∫Ωu(1−u−v1+au+bv)dx=0. |
Thus, there must exists a point x2∈Ω such that
1−u(x2)−v(x2)1+au(x2)+bv(x2)=0. |
Furthermore, we get
max¯Ωv≥v(x2)≥v(x2)1+au(x2)+bv(x2)=1−u(x2)>0. |
According to the inequality (2.3), if min¯Ωv→0, we have max¯Ωv→0. Hence, we get a contradiction. This shows v has a positive lower bound. This completes the proof.
Remark 2.4. Theorem 2.3 shows that if d1 is not too large or not too small, d2 is not too small and d3 is not too large, then the solutions of system (1.2) are bounded, that is to say, there exists a ball B(C) such that all the positive solution (u,v) of (1.2) satisfying (u,v)∈B(C). We can also conclude that the bound of the solution is not constrained by the cross diffusion d4.
The purpose of this section is to study the non-existence and existence of non-constant positive stationary solutions of system (1.2) by taking the self- and cross-diffusions as parameters. The main method used to prove the existence of non-constant positive solutions is the Leray-Schauder degree. The results show that the the self-diffusion d2 and the fractional type cross-diffusion d4 can create spatial patterns while d1 and d3 failed.
This subsection is devoted to investigating the non-existence of non-constant positive solutions of system (1.2). We mainly use the energy integral method and the well-known Poincaré inequality. Let μ1 be the smallest positive eigenvalue of the operator −Δ subject to the homogeneous Neumann boundary condition. For convenience, we set
˜u=1|Ω|∫Ωudxand˜v=1|Ω|∫Ωvdx. |
Through some calculations and analysis, we obtain the following result.
Theorem 3.1. Let d1,d2,d4 be fixed positive constants and d3=0. If there exists a positive constant ˜C=˜C(Θ,D1,Ω), such that d1>max{D1,˜C(1+d22d24)},d2>D2 and d4<D4, then system (1.2) has no non-constant positive solution.
Proof. Assume that (u,v) is a positive solution of system (1.2). We multiply u−˜uu and v−˜vv to the equations of system (1.2), respectively, and integrate the equation by parts in Ω. Then, we have
d1∫Ω˜uu2(1+d3v)|▽u|2dx+d2∫Ω˜vv2(1+d41+u)|▽v|2dx−d2d4∫Ω˜vv(1+u)2▽u▽vdx=∫Ω[−1+a˜v(1+a˜u+b˜v)(1+au+bv)](u−˜u)2dx−∫Ω[bc˜u(1+a˜u+b˜v)(1+au+bv)](v−˜v)2dx+∫Ω[b˜v−(1+a˜u+b˜v)(1+a˜u+b˜v)(1+au+bv)+c(1+a˜u+b˜v)−ac˜u(1+a˜u+b˜v)(1+au+bv)](u−˜u)(v−˜v)dx. |
According to Theorem 2.3, for d1 with a fixed small D1, there exists a large enough positive constant C1=C1(Θ,D1,D2,D4,Ω), such that
C1−1<u(x),v(x),˜u,˜v<C1. |
Therefore, we have
d1C31∫Ω|▽u|2dx+d2C31∫Ω|▽v|2dx≤∫ΩaC1(u−˜u)2dx−∫Ωbc(1+aC1+bC1)2C1(v−˜v)2dx+∫Ω(bcC1+c)|u−˜u||v−˜v|dx+d2d4C21∫Ω|▽u||▽v|dx. |
It follows the Young's inequality[19,20,21],
d1C31∫Ω|▽u|2dx+d2C31∫Ω|▽v|2dx≤∫Ω[aC1+(bcC1+c)24K](u−˜u)2dx+∫Ω[K−bc(1+aC1+bC1)2C1](v−˜v)2dx+d22d24C414K∫Ω|▽u|2dx+K∫Ω|▽v|2dx, |
where K is a arbitrary small positive constant. For convenience, we take d3=0 and K=min{d2C31,bc(1+aC1+bC1)2C1}. By employing the Cauchy inequality and Poincaré inequality[22,23], we have
d1C31∫Ω|▽u|2dx≤∫Ω[aC1+(bcC1+c)24K](u−˜u)2dx+d22d24C414K∫Ω|▽u|2dx≤C2(Θ,D1,D2,D4,Ω)∫Ω(u−˜u)2dx+d22d24C3(Θ,D1,D2,D4,Ω)∫Ω|▽u|2dx≤(C2(Θ,D1,D2,D4,Ω)μ1+d22d24C3(Θ,D1,D2,D4,Ω))∫Ω|▽u|2dx≤C4(Θ,D1,D2,D4,Ω)(1+d22d24)∫Ω|▽u|2dx. |
Therefore, we can assert that u≡˜u,v≡˜v if d1>max{D1,˜C(1+d22d24)},d2>D2 and d4<D4.
In this subsection, we mainly consider the existence of non-constant positive solution of system (1.2) by taking the self- and cross- diffusion coefficients as parameters. Particularly, combing with Theorems 2.3 and 3.1, we consider the cases that the self-diffusion d2 or the cross-diffusion d4 is large enough. The key method used in this subsection to prove the existence of non-constant positive solutions is the well-known Leray-Schauder degree theory [24,25,26], which has been extensively used in many different papers. In order to establish the existence of stationary patterns of system (1.2), for convenience, we first introduce some notations and definitions. We define
ω=(u,v)T, G(ω)=(u(1−u−v1+au+bv),v(−d+cu1+au+bv))T, |
Φ(ω)=(ϕ1(ω),ϕ2(ω))T=(d1(1+d3v)u,d2(1+d41+u)v)T, Λ=(d1,d2,d3,d4) |
and set
X={ω=(u,v)T∈(C2(Ω)∩C1(¯Ω))2|∂νu=∂νv=0 on ∂Ω}, |
X+={ω∈X|u>0,v>0 on ¯Ω}. |
B(C)={(u,v)T∈X|C−1<u,v<C}, |
where C is a positive constant provided by Theorem 2.3. Let 0=μ0<μ1<μ2<⋯ be the eigenvalues of the operator −Δ and {μi,ψi}∞i=0 be a complete set of eigenpairs for the operator −Δ in Ω under homogeneous Neumann boundary condition. Moreover, we can decompose X=⊕∞i=0Xi and Xi=⊕dimE(μi)j=1Xij, where Xi is the eigenspace corresponding to the eigenvalue μi.
Therefore, system (1.2) can be rewritten as
{−ΔΦ(ω)=G(ω),x∈Ω,∂νω=0,x∈∂Ω. | (3.1) |
It is clear that system (3.1), as well as system (1.2), has a constant positive equilibrium point, denoted by ω∗=(u∗, v∗)T, where u∗, v∗ are given in Section 2.
By direct computation, we obtain ∂Φ(u,v)∂(u,v)=[d1+d1d3vd1d3u−d2d4v(1+u)2d2+d2d41+u] and it is easy to show that det[∂Φ(u,v)∂(u,v)]>0 for all non-negative solutions (u,v)T. Therefore, we know that Φ−1ω exists and det[∂Φ(u,v)∂(u,v)]−1 is positive. Then, system (1.2) can also be rewritten as
F(Λ;ω)≜ω−(I−Δ)−1{Φ−1ω(ω)[G(ω)+∇ωΦωω(ω)∇ω]+ω}=0, | (3.2) |
where I is the identity operator and (I−Δ)−1 is the inverse of the operator I−Δ on X with the homogeneous Neumann boundary condition. Since F(Λ;⋅) is a compact perturbation of the identity operator, the Leray-Schauder degree deg(F(Λ;⋅),0,B) is well-defined if F(Λ;ω)≠0 for all ω∈∂B. Furthermore, we notice that the linearizition of the operator F(Λ;ω) at the equilibrium point ω∗ is
DωF(Λ;ω∗)=I−(I−Δ)−1[Φ−1ω(ω∗)Gω(ω∗)+I], |
and Xi is invariant under DωF(Λ;ω∗) for every integer i≥0. What's more, one can check that λ is an eigenvalue of the operator DωF(Λ;ω∗) on Xi if and only if λ is an eigenvalue of the matrix
I−11+μi[Φ−1ω(ω∗)Gω(ω∗)+I]=11+μi[μiI−Φ−1ω(ω∗)Gω(ω∗)]. |
Denote
H(Λ,ω∗;μi)≜det[μiI−Φ−1ω(ω∗)Gω(ω∗)]=det{[Φω(ω∗)]−1}det[μΦω(ω∗)−Gω(ω∗)]. |
As the sign of the
det{I−11+μi[Φ−1ω(ω∗)Gω(ω∗)+I]} |
is determined by the number of negative eigenvalue of the matrix
I−11+μi[Φ−1ω(ω∗)Gω(ω∗)+I], |
then both H(Λ,ω∗;μi) and det{I−11+μi[Φ−1ω(ω∗)Gω(ω∗)+I]} have the same sign. Hence, if H(Λ,ω∗;μi)≠0, the number of eigenvalues with negative real parts of DωF(Λ;ω∗) on Xi is odd if and only if H(Λ,ω∗;λi)<0.
If H(Λ,ω∗;μi)≠0 for all integer i≥0, then 0 is not an eigenvalue of the operator DωF(Λ;ω∗). This indicates that DωF(Λ;ω∗) is a homeomorphism operator from the space X to X. Then the implicit function theorem shows that the equilibrium point ω=ω∗ is an isolated solution of equation F(Λ;ω)=0. In summary, according to Leray-Schauder degree theory, we present the following results (One can refer to [27,28]).
Lemma 3.3. Assume that the matrix μiI−Φ−1ω(ω∗)Gω(ω∗) is non-singular for each i>0. Then
index(F(Λ;⋅),ω∗)=(−1)τ,τ=∑i≥0,H(D,ω∗;μi)<0dimE(μi). |
In order to calculate index(F(Λ;⋅),ω∗), we will consider the sign of H(Λ,ω∗;μi) in detail. Notice that det{Φω(ω∗)−1}>0, so we only need to consider det[μΦω(ω∗)−Gω(ω∗)]. Review that
{u∗=12bc[ad+bc−c+√(ad+bc−c)2+4bcd],v∗=cd(1−u∗)u∗ |
and
1−u∗−v∗1+au∗+bv∗=0, −d+cu∗1+au∗+bv∗=0. |
Then, by direct calculation, we obtain
Φω(ω∗)=[d1(1+d3v∗)d1d3u∗−d2d4v∗(1+u∗)2d2+d2d41+u∗], |
Gω(ω∗)=[1−2u∗−v∗(1+bv∗)(1+au∗+bv∗)2−u∗(1+au∗)(1+au∗+bv∗)2cv∗(1+bv∗)(1+au∗+bv∗)2−bcu∗v∗(1+au∗+bv∗)2] |
=[1−2u∗−(1−u∗)2(b+1v∗)−d2c2(a+1u∗)c(1−u∗)2(b+1v∗)bd(u∗−1)] |
and
det[μΦω(ω∗)−Gω(ω∗)]=Aμ2+Bμ+C≜ψ(μ), | (3.3) |
where
A=d1(1+d3v∗)(d2+d2d41+u∗)+d1d2d3d4u∗v∗(1+u∗)2>0,B=(1−u∗)[d1d3du∗(1+bv∗)+d2d4c(1−u∗)(1+bv∗)(1+u∗)2+d1bd(1+d3v∗)−(d2+d2d41+u∗)c(1+bv∗)1+au∗+bv∗],C=d2c2(a+1u∗)c(1−u∗)2(b+1v∗)−[1−2u∗−(1−u∗)2(b+1v∗)]bd(1−u∗). |
Denote μ1(Λ) and μ2(Λ) be the two roots of the equation Aμ2+Bμ+C=0 with Reμ1(Λ)≤Reμ2(Λ). We notice that C>0 guaranteed by 2u∗+(1−u∗)2(b+1v∗)>1, and we obtain
μ1(Λ)μ2(Λ)=CA>0. |
Moreover, we have
limd2→+∞ψ(μ)d2=A1μ2+B1μ, limd4→+∞ψ(μ)d4=A2μ2+B2μ, |
where
A1=limd2→+∞Ad2=d1(1+d3v∗)(1+d41+u∗)+d1d3d4u∗v∗(1+u∗)2>0,B1=limd2→+∞Bd2=(1−u∗)[d4c(1−u∗)(1+bv∗)(1+u∗)2−(1+d41+u∗)c(1+bv∗)1+au∗+bv∗],A2=limd4→+∞Ad4=d1d2(1+d3v∗)11+u∗+d1d2d3u∗v∗(1+u∗)2>0,B2=limd4→+∞Bd4=B=(1−u∗)[d2c(1−u∗)(1+bv∗)(1+u∗)2−d21+u∗c(1+bv∗)1+au∗+bv∗]. |
We notice that A1,A2 are always positive while B1,B2 may change the sign. In conclusion, we can obtain the following lemmas.
Lemma 3.4. Assume that di(i=1,3,4) are all fixed and B1<0. Then there exists a positive constant ˆd2, such that for d2≥ˆd2, the two roots ˜μ1(d2) and ˜μ2(d2) of the equation ψ(μ)=0 are all real and satisfy
limd2→+∞˜μ1(d2)=0, limd2→+∞˜μ2(d2)=−B1A1, |
where
{ψ(μ,d2)<0, if μ∈(˜μ1(d2),˜μ2(d2)),ψ(μ,d2)>0, if μ∈(−∞,˜μ1(d2))∪(˜μ2(d2),+∞). | (3.4) |
Lemma 3.5. Assume that di(i=1,2,3) are all fixed and B2<0. Then there exists a positive constant ˆd4, such that for d4≥ˆd4, the two roots ˜μ1(d4) and ˜μ2(d4) of the equation ψ(μ)=0 are all real and satisfy
limd4→+∞˜μ1(d4)=0, limd4→+∞˜μ2(d4)=−B2A2, |
where
{ψ(μ,d4)<0, if μ∈(˜μ1(d4),˜μ2(d4)),ψ(μ,d4)>0, if μ∈(−∞,˜μ1(d4))∪(˜μ2(d4),+∞). | (3.5) |
In the following, by using Leray-Schauder degree theory [29,30], we investigate the existence of non-constant positive solutions to system (1.2) with respect to the diffusion coefficients di,i=1,2,3,4. We mainly consider the cases that d2 or d4 is large enough and in view of Lemmas 3.4 and 3.5, we can obtain the following theorems. We only prove Theorems 3.6 and 3.7 can be finished similarly.
Theorem 3.6. Assume that di(i=1,2,3) are all fixed and B1<0. For ˜μ be given by the limit of Lemma 3.5, if ˜μ∈(μn,μn+1) for some integer n≥1 and the sum τn=∑ni=1dimE(μi) is odd, then there exists a positive constant ˆd4 such that system (1.2) has at least one non-constant positive solution when d4≥ˆd4.
Proof. According to Lemma 3.4, there exists a positive constant ˆd4, such that (3.4) holds and 0=μ0<˜μ1<˜μ2, ˜μ2∈(μn,μn+1) when d4≥ˆd4.
We shall prove the result by contradiction. Suppose on the contrary that the result is not true for some d4=˜d4≥ˆd4, that is, system (1.2) does not have any positive non-constant positive solution and index(F(Λ;⋅),ω∗)=1 when d4≥ˆd4. If we take Λ∗=(ˆd1,d2,d3,ˆd4), then,
deg(F(Λ∗;⋅),0,B)=1, | (3.6) |
where, ˆd1 is a moderately large constant provided by Theorem 2.3.
For t∈[0,1], we define a homotopy as
{ −Δ[(td1+(1−t)ˆd1+td1d3v)u]=u(1−u−v1+au+bv),x∈Ω,−Δ[(td2+((1−t)ˆd4+td4)d21+u)v]=v(−d+cu1+au+bv),x∈Ω,∂νu=∂νv=0,x∈∂Ω. | (3.7) |
Thus, ω is a non-constant positive solution of system (3.7) if and only if ω is a non-constant positive solution of the following problem
˜F(t,Λ;ω)≜ω−(I−Δ)−1{˜Φω}−1(ω,t)[G(ω)+∇ω˜Φωω(ω,t)∇ω]+ω}=0, | (3.8) |
on X, where ˜Φ=((td1+(1−t)ˆd1+td1d3v)u,(td2+((1−t)ˆd4+td4)d21+u)v). It is clear that
Φ(ω)=˜Φ(ω,1), F(Λ;ω)=˜F(1,Λ;ω), F(Λ∗;ω)=˜F(0,Λ;ω). |
Theorem 3.1 shows that ˜F(0,Λ;ω)=0 only has the constant positive solution ω∗ in X. Through some calculation, we obtain
Dω˜F(t,ω∗)=I−(I−Δ)−1[Φ−1ω(t,ω∗)Gω(ω∗)+I]. |
Furthermore, in view of Lemma 3.5, for t=1, we have
{H(Λ,ω∗;μi)<0, when 1≤i≤n,H(Λ,ω∗;μi)>0, when i>n. |
Then, 0 is not an eigenvalue of the matrix μiI−[Φω]−1Gω(ω∗), and τn=∑ni=1dimE(μi) is odd. Moreover, in view of Lemma 3.3, we know that
index(˜F(1,Λ;⋅),ω∗)=(−1)τ=−1. |
According to Theorem 2.3, we know that all the positive solutions (u,v) of the system (1.2) are in B(C) for large enough constant C. Therefore, system (3.6) has no solution on ∂B for any t∈[0,1], and deg(˜F(t,Λ;⋅),0,B) is well defined. According to the homotopy invariance of the Leray-Schauder degree, we obtain
deg(˜F(0,Λ;⋅),0,B)=deg(˜F(1,Λ;⋅),0,B). | (3.9) |
Therefore, we have
deg(˜F(0,Λ;⋅),0,B)=deg(F(Λ∗;⋅),0,B)=index(F(Λ;⋅),ω∗)=1, | (3.10) |
deg(˜F(1,Λ;⋅),0,B)=deg(F(Λ;⋅),0,B)=index(˜F(1,Λ;⋅),ω∗)=−1. | (3.11) |
From (3.9)–(3.11), we obtain a contradiction. Hence, system (1.2) has at least one non-constant positive solution and the proof is finished.
Theorem 3.7. Assume that di(i=1,3,4) are all fixed and B2<0. For ˜μ′ be given by the limit of Lemma 3.4, if ˜μ′∈(μn,μn+1) for some integer n≥1 and the sum τn=∑ni=1dimE(μi) is odd, then there exists a positive constant ˆd2 such that system (1.2) has at least one non-constant positive solution when d2≥ˆd2.
This paper investigates the existence of positive stationary solutions of a predator-prey system with Beddington-DeAngelis functional response and fractional cross-diffusion d4 subjected to the homogeneous Neumann boundary condition. The priori estimate result shows that if d1 is not too large or not too small, d2 is not too small and d3 is not too large, then the solutions of system (1.2) are bounded and we can assert that the bound of the solution is not constrained by the cross diffusion d4. Moreover, from the proof the non-existence of non-constant positive solution, we obtain the sufficient condition for the non-existence, that is, d2d4 is small enough. Finally, we discuss the existence of non-constant positive solution and the results indicate that the system admits a non-constant positive solution provided by the self-diffusion d2 or the cross-diffusion d4 is large enough, which means the diffusion d2 and the fractional type cross-diffusion d4 can create spatial patterns.
Furthermore, this study contributes to the growing field of fractional diffusion models in ecology. Fractional diffusion is a generalization of classical diffusion that allows for non-local interactions and has been shown to better capture the long-range effects observed in many ecological systems. The results of this paper demonstrate the potential for fractional cross-diffusion to play an important role in determining the existence and properties of positive solutions in predator-prey systems.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We would like to give our sincere thanks to the anonymous referees and the handing editors for their valuable comments and helpful suggestions on this manuscript.
This work is supported in part by the Basic Research Program of Natural Science of Shaanxi Province (No. 2022JM-055), and also by the Science Research Program of Xi'an Eurasia University (No. 2022GCPY01).
No conflict of interest exits in the submission of this manuscript, and manuscript is approved by all authors for publication. I would like to declare on behalf of my co-authors that the work described was original research that has not been published previously, and not under consideration for publication elsewhere, in whole or in part. All the authors listed have approved the manuscript that is enclosed.
[1] |
W. Chen, M. Wang, Qualitative analysis of predator-prey models with Beddington-DeAngelis functional response and diffusion, Math. Comput. Model., 42 (2005), 31–44. http://dx.doi.org/10.1016/j.na.2006.08.022 doi: 10.1016/j.na.2006.08.022
![]() |
[2] |
R. S. Cantrell, C. Cosner, On the dynamics of predator-prey models with the Beddington-DeAngelis functional response, J. Math. Anal. Appl., 257 (2001), 206–222. https://doi.org/10.1006/jmaa.2000.7343 doi: 10.1006/jmaa.2000.7343
![]() |
[3] |
D. T. Dimitrov, H. V. Kojouharov, Complete mathematical analysis of predator-prey models with linear prey growth and Beddington-DeAngelis, Appl. Math. Comput., 162 (2005), 523–538. https://doi.org/10.1016/j.amc.2003.12.106 doi: 10.1016/j.amc.2003.12.106
![]() |
[4] |
D. Luo, Q. Wang, Global bifurcation and pattern formation for a reaction-diffusion predator-prey model with prey-taxis and double Beddington-DeAngelis functional responses, Nonlinear Anal.-Real, 67 (2022), 103638. https://doi.org/10.1016/j.nonrwa.2022.103638 doi: 10.1016/j.nonrwa.2022.103638
![]() |
[5] |
Y. Lou, W. M. Ni, Diffusion, self-diffusion and cross-diffusion, J. Differ. Equations, 131 (1996), 79–131. https://doi.org/10.1006/jdeq.1996.0157 doi: 10.1006/jdeq.1996.0157
![]() |
[6] |
Y. Lou, W. M. 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] | F. Yi, Turing instability of the periodic solutions for reaction-diffusion systems with cross-diffusion and the patch model with cross-diffusion-like coupling, 281 (2021), 379–410. https://doi.org/10.1016/j.jde.2021.02.006 |
[8] |
Q. Meng, L. Yang, Steady state in a cross-diffusion predator-prey model with the Beddington-DeAngelis functional response, Nonlinear Anal.-Real, 45 (2019), 401–413. https://doi.org/10.1016/j.nonrwa.2018.07.012 doi: 10.1016/j.nonrwa.2018.07.012
![]() |
[9] |
J. F. Zhang, Spatial patterns of a fractional type cross-diffusion Holling-Tanner model, Comput. Math. Appl., 76 (2018), 957–965. https://doi.org/10.1016/j.camwa.2018.05.033 doi: 10.1016/j.camwa.2018.05.033
![]() |
[10] |
P. Xue, Y. Jia, C. Ren, X. Li, Non-constant positive solutions of a general Gause-type predator-prey system with self- and cross- diffusions, Math. Model. Nat. Pheno., 16 (2021), 1–25. https://doi.org/10.1051/mmnp/2021017 doi: 10.1051/mmnp/2021017
![]() |
[11] |
Y. Jia, P. Xue, Effects of the self- and cross-diffusion on positive steady states for a generalized predator-prey system, Nonlinear Anal.-Real, 32 (2016), 229–241. https://doi.org/10.1016/j.nonrwa.2016.04.012 doi: 10.1016/j.nonrwa.2016.04.012
![]() |
[12] |
C. Li, Existence of positive solution for a cross-diffusion predator-prey system with Holling type-II functional response, Chaos Soliton. Fract., 99 (2017), 226–232. https://doi.org/10.1016/j.chaos.2017.04.001 doi: 10.1016/j.chaos.2017.04.001
![]() |
[13] |
S. Li, Y. Yamada, Effect of cross-diffusion prey-preydator model with a protection zone II, J. Math. Anal. Appl., 461 (2018), 971–992. https://doi.org/10.1016/j.jmaa.2017.12.029 doi: 10.1016/j.jmaa.2017.12.029
![]() |
[14] | A. Okubo, Diffusion and ecological problems: Mathematical models, Springer-Verlag, Berlin, 1980. |
[15] |
J. Cao, H. Sun, P. Hao, P. Wang, Bifurcation and turing instability for a predator-prey model with nonlinear reaction cross-diffusion, Appl. Math. Model., 89 (2021), 1663–1677. https://doi.org/10.1016/j.apm.2020.08.030 doi: 10.1016/j.apm.2020.08.030
![]() |
[16] |
X. Chen, Y. Qi, M. Wang, A strongly coupled predator-prey system with non-monotonic functional response, Nonlinear Anal., 67 (2007), 1966–1979. https://doi.org/10.1016/j.na.2006.08.022 doi: 10.1016/j.na.2006.08.022
![]() |
[17] | Y. Lou, W. M. Ni, S. Yotsutani, On a limiting system in the Lotka-Volterra competition with cross-diffusion, Discrete Cont. Dyn. Syst., 10 (2004), 435–458. |
[18] |
J. H. P. Dawes, M. O. Souza, A derivation of Holling's type I, II and III functional responses in predator-prey systems, J. Theor. Biol., 327 (2013), 11–22. https://doi.org/10.1016/j.jtbi.2013.02.017 doi: 10.1016/j.jtbi.2013.02.017
![]() |
[19] | D. Gilbarg, N. S. Trudinger, Elliptic partial differential equation of second order, Springer-Verlag, Berlin, 2001. |
[20] | X. Zeng, Non-constant positive steady states of a prey-predator system with cross-diffusions, J. Math. Anal. Appl., 332 (2007), 989–1009. https://doi.org/10.1006/jmaa.2000.7343 |
[21] |
R. K. Ghaziani, W. Govaerts, C. Sonck, Resonance and bifurcation in a discrete-time predator-prey system with Holling functional response, Nonlinear Anal.-Real, 13 (2012), 1451–1465. https://doi.org/10.1016/j.nonrwa.2011.11.009 doi: 10.1016/j.nonrwa.2011.11.009
![]() |
[22] |
R. Peng, J. P. 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. https://doi.org/10.1016/j.jde.2009.03.008 doi: 10.1016/j.jde.2009.03.008
![]() |
[23] |
Y. Jia, A sufficient and necessary condition for the existence of positive solutions for a prey-predator system with Ivlev-type functional response, Appl. Math. Lett., 24 (2011), 1084–1088. https://doi.org/10.1016/j.aml.2011.01.027 doi: 10.1016/j.aml.2011.01.027
![]() |
[24] |
J. Sugie, Two-parameter bifurcation in a predator-prey system of Ivlev type, J. Math. Anal. Appl., 217 (1998), 349–371. https://doi.org/10.1006/jmaa.1997.5700 doi: 10.1006/jmaa.1997.5700
![]() |
[25] |
Y. Jia, J. Wu, H. K. Xu, Positive solutions of a Lotka-Volterra competition model with cross-diffusion, Comput. Math. Appl., 68 (2014), 1220–1228. https://doi.org/10.1016/j.camwa.2014.08.016 doi: 10.1016/j.camwa.2014.08.016
![]() |
[26] |
W. Ko, K. Ryu, A qualitative study on general Gause-type predator-prey models with constant diffusion rates, J. Math. Anal. Appl., 344 (2008), 217–230. https://doi.org/10.1016/j.jmaa.2008.03.006 doi: 10.1016/j.jmaa.2008.03.006
![]() |
[27] |
K. Oeda, Effect of cross-diffusion on the stationary problem of a prey-predator model with a protection zone, J. Differ. Equations, 250 (2011), 3988–4009. https://doi.org/10.1016/j.jde.2011.01.026 doi: 10.1016/j.jde.2011.01.026
![]() |
[28] |
Y. Peng, T. Zhang, Turing instability and pattern induced by cross-diffusion in a predator-prey system with Allee effect, Appl. Math. Comput., 275 (2016), 1–12. https://doi.org/10.1016/j.amc.2015.11.067 doi: 10.1016/j.amc.2015.11.067
![]() |
[29] | L. Nirenberg, Topics in nonlinear functional analysis, American Mathematical Socity, Providence, Rhode Island, 2001. |
[30] |
H. Wang, P. Liu, Pattern dynamics of a predator-prey system with cross-diffusion, Allee effect and generalized Holling IV functional response, Chaos Soliton. Fract., 171 (2023), 113456. https://doi.org/10.1016/j.chaos.2023.113456 doi: 10.1016/j.chaos.2023.113456
![]() |