Processing math: 100%
Research article Special Issues

Predator-prey systems with defense switching and density-suppressed dispersal strategy


  • In this paper, we consider the following predator-prey system with defense switching mechanism and density-suppressed dispersal strategy

    {ut=Δ(d1(w)u)+β1uvwu+vα1u,xΩ,t>0,vt=Δ(d2(w)v)+β2uvwu+vα2v,xΩ,t>0,wt=Δwβ3uvwu+v+σw(1wK),xΩ,t>0,uν=vν=wν=0,xΩ,t>0,(u,v,w)(x,0)=(u0,v0,w0)(x),xΩ, 

    where ΩR2 is a bounded domain with smooth boundary. Based on the method of energy estimates and Moser iteration, we establish the existence of global classical solutions with uniform-in-time boundedness. We further prove the global stability of co-existence equilibrium by using the Lyapunov functionals and LaSalle's invariant principle. Finally we conduct linear stability analysis and perform numerical simulations to illustrate that the density-suppressed dispersal may trigger the pattern formation.

    Citation: Jiawei Chu, Hai-Yang Jin. Predator-prey systems with defense switching and density-suppressed dispersal strategy[J]. Mathematical Biosciences and Engineering, 2022, 19(12): 12472-12499. doi: 10.3934/mbe.2022582

    Related Papers:

    [1] Sourav Kumar Sasmal, Jeet Banerjee, Yasuhiro Takeuchi . Dynamics and spatio-temporal patterns in a prey–predator system with aposematic prey. Mathematical Biosciences and Engineering, 2019, 16(5): 3864-3884. doi: 10.3934/mbe.2019191
    [2] 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
    [3] 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
    [4] 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
    [5] Raimund Bürger, Gerardo Chowell, Elvis Gavilán, Pep Mulet, Luis M. Villada . Numerical solution of a spatio-temporal predator-prey model with infected prey. Mathematical Biosciences and Engineering, 2019, 16(1): 438-473. doi: 10.3934/mbe.2019021
    [6] 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
    [7] 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
    [8] Xiaoying Wang, Xingfu Zou . Pattern formation of a predator-prey model with the cost of anti-predator behaviors. Mathematical Biosciences and Engineering, 2018, 15(3): 775-805. doi: 10.3934/mbe.2018035
    [9] Jin Zhong, Yue Xia, Lijuan Chen, Fengde Chen . Dynamical analysis of a predator-prey system with fear-induced dispersal between patches. Mathematical Biosciences and Engineering, 2025, 22(5): 1159-1184. doi: 10.3934/mbe.2025042
    [10] Yansong Pei, Bing Liu, Haokun Qi . Extinction and stationary distribution of stochastic predator-prey model with group defense behavior. Mathematical Biosciences and Engineering, 2022, 19(12): 13062-13078. doi: 10.3934/mbe.2022610
  • In this paper, we consider the following predator-prey system with defense switching mechanism and density-suppressed dispersal strategy

    {ut=Δ(d1(w)u)+β1uvwu+vα1u,xΩ,t>0,vt=Δ(d2(w)v)+β2uvwu+vα2v,xΩ,t>0,wt=Δwβ3uvwu+v+σw(1wK),xΩ,t>0,uν=vν=wν=0,xΩ,t>0,(u,v,w)(x,0)=(u0,v0,w0)(x),xΩ, 

    where ΩR2 is a bounded domain with smooth boundary. Based on the method of energy estimates and Moser iteration, we establish the existence of global classical solutions with uniform-in-time boundedness. We further prove the global stability of co-existence equilibrium by using the Lyapunov functionals and LaSalle's invariant principle. Finally we conduct linear stability analysis and perform numerical simulations to illustrate that the density-suppressed dispersal may trigger the pattern formation.



    Defense switching means that prey species pay more attention on guarding against the relatively more abundant population predator [1], certain fish species in Lake Tanganyika against two phenotypes (dextral and sinistral) of cichlid Perissodus microlepis [2] is a typical example. The dextral and sinistral phenotypes attack the prey fishes from the left-side and right-side, respectively. Pretend that the population of dextral individuals is more abundant, then prey fishes tend to be more defensive against the attacks from left-side, which leads to greater hunting success for sinistral individuals (relatively rare population). Based on a simple Lotka-Volterra equations of a two-predator one-prey system, Saleem et al. proposed a defensive switching model [1]

    {ut=u(α1+f1w),vt=v(α2+f2w),wt=w(σf1uf2v), (1.1)

    with the predatory rates functions

    {f1(u,v)=β11+uv,f2(u,v)=β2(1f1β1), (1.2)

    where u:=u(x,t) and v:=v(x,t) denote the density of two predators while w:=w(x,t) is the prey density. The parameters αj(j=1,2) account the death rate, βj>0(j=1,2) are the predation coefficients and σ>0 accounts for the growth rate of the prey species. Moreover, the predatory rates f1 and f2, also called "defensive switching functions", possess a characteristic property that the rate of the prey attacked by a predator will decrease if this predator population becomes much more abundant than the population of another predator. Specifically, when a predator population becomes large, the prey species guards against it more vigilantly and switches to another predator, which is in relatively small population, to keep its individual from being hunted too much. Such prey behaviors result in less successful hunting for abundant population predator species and more successful hunting for relatively rare one [1].

    In light of the defensive switching model Eqs (1.1) and (1.2) in [1], Pang and Wang [3] considered the following reaction-diffusion system by introducing the random movements of species and the intra-specific interaction between the prey in Eq (1.1)

    {ut=d1Δu+β1uvwu+vα1u,vt=d2Δv+β2uvwu+vα2v,wt=d3Δw(β1+β2)uvwu+v+σw(1wK), (1.3)

    where constants dj>0(j=1,2,3) account for diffusion coefficients and the positive parameter K represents the environmental carrying capacity to the prey species. We note that the interaction mechanism between the predators and prey in the defensive switching model is substantially different from that in the ratio-dependent predator-prey system [4] though they seem to have some similar structures.

    In [1], Saleem et al. proved that the co-existence steady state is globally asymptotically stable except the case where the two predators have the same mortality rates; Otherwise, the system has a periodic solution. Pang and Wang [3] proved the co-existence steady states is globally asymptotically stable no matter α1 and α2 are equal or not and hence no pattern formation will arise from the system (1.3). If the term d1Δu is replaced by a cross diffusion term Δ(d1u+kuϵ+v2)(k,ϵ>0), however, they showed that the cross diffusion between predators can drive stationary patterns by using Leray-Schauder degree theory. Recently, the system (1.3) with prey-taxis (the cross diffusion between the predators and prey) was considered by Wang and Guo [5]. They established the existence of globally bounded solutions and global stability of constant steady states for small prey-tactic coefficient and numerically demonstrated that strong prey-taxis can induce the pattern formation if the two predators have different mortality rates. Subsequently, the existence of nonconstant steady states was obtained for some range of repulsive prey-taxis coefficient with more general functional response functions in [6]. From the above results, we find that the dispersal strategies (like cross-diffusion or prey-taxis) play important roles in determining the population distribution profiles.

    In this paper, we shall consider a different dispersal strategy - density-suppressed diffusion, which was first used in [7] to describe the directed movement of predators in the predator-prey systems to fit the experimental observations. This type of diffusion assumes that the predator's diffusion decreasingly depends on the density distribution of the prey. As shown in [8], the density-suppressed diffusion can explain the heterogenous population distribution observed in the field experiment of [7] while random diffusion can not. Hence density-suppressed diffusion is a dispersal strategy employed in the predator-prey system. By taking into account the density-suppressed diffusion, the defensive switching model reads

    {ut=Δ(d1(w)u)+β1uvwu+vα1u,vt=Δ(d2(w)v)+β2uvwu+vα2v,wt=d3Δw(β1+β2)uvwu+v+σw(1wK), (1.4)

    where dj(w)>0 and dj(w)<0(j=1,2). Note that the property dj(w)<0 means that predators will decrease their random diffusion rates at higher density of the prey species in order for predation. If we expand the density-suppressed diffusion

    Δ(dj(w)ϕ)=(dj(w)ϕ+ϕdj(w)w),  ϕ=u,v,

    we find that the density-suppressed diffusion indeed intrinsically includes both random diffusion and advection (prey-taxis) components. The differences from the spatial models considered in [3,5,6] is that here both diffusion and advection coefficient are not constant but functions of the prey density. It was shown that the density-suppressed diffusion in the predator-prey systems may generate spatially non-homogeneous patterns that the random diffusion can not do [8] and may bring substantially different dynamics [9].

    Beyond the predator-prey systems, the density-suppressed diffusion has already been commonly used in the modeling of other biological processes such as the chemotaxis [10,11], bacterial movement [12,13] and so on. Since the possible degeneracy caused by the density-suppressed diffusion brings considerable challenges for analysis, the studies of these biological models with density-suppressed diffusion have been increasingly attracting attentions and produced many interesting analytical results [14,15,16,17,18,19,20,21,22,23,24,25,26,27,28] alongside rich numerical simulations demonstrating complex dynamics and patterns [29,30,31].

    The purpose of this paper is to study the global dynamics of the defensive switching model with density-suppressed diffusion including global existence and asymptotic behavior of solutions as well as pattern formations. Specifically we consider the following problem

    {ut=Δ(d1(w)u)+β1uvwu+vα1u,xΩ,t>0,vt=Δ(d2(w)v)+β2uvwu+vα2v,xΩ,t>0,wt=Δwβ3uvwu+v+σw(1wK),xΩ,t>0,uν=vν=wν=0,xΩ,t>0,(u,v,w)(x,0)=(u0,v0,w0)(x),xΩ, (1.5)

    where ΩR2 is a bounded domain with smooth boundary. The parameters αj(j=1,2),βj(j=1,2,3),σ and K are all positive constants. Note that we have assumed d3=1 without loss of generality and consider more general constant β3 compared to Eq (1.4). We suppose the density-suppressed diffusion coefficients dj(w) satisfy the following conditions:

    (H0) dj(w)C2([0,)) with dj(w)>0 and dj(w)0 for all w0,  j=1,2.

    We further suppose the initial data satisfy

    (u0,v0,w0)[W1,(Ω)]3withu0,v0>0,w00. (1.6)

    Then our first result on the global existence of solutions is stated in the following theorem.

    Theorem 1.1 (Global boundedness). Let ΩR2 be a bounded domain with smooth boundary. Assume the assumption (H0) holds and the initial data (u0,v0,w0) satisfy Eq (1.6). Then there exists a uniquely determined triple (u,v,w) of nonnegative functions which solves Eq (1.5) classically in Ω×(0,) and satisfies

    u(,t)L+v(,t)L+w(,t)W1,M0  for all t>0,

    where M0>0 is a constant independent of t. Particularly, one has

    0<w(x,t)K:=max{w0L,K}  for all  (x,t)Ω×(0,).

    Remark 1.2. Note that in works [5,6] considering the defensive switching model with prey-taxis, the smallness of prey-taxis coefficient was required to ensure the global boundedness of solutions. Here we obtain the existence of global classical solutions with uniform-in-time boundedness without any smallness assumptions on the parameters.

    Next, we shall show the global stability of constant steady states. In fact it is straightforward to find that a positive constant steady state (u,v,w) exists if and only if K>α1β1+α2β2, where u,v,w>0, satisfying

    α1=β1vwu+v,  α2=β2uwu+v,  σ(1wK)=β3uvu+v, (1.7)

    which can be solved to obtain

    u=σβ3(1wK)(1+α2β1α1β2), v=σβ3(1wK)(1+α1β2α2β1), w=α1β1+α2β2. (1.8)

    Then we can show that (u,v,w) is globally asymptotically stable under certain conditions, as stated in the following theorem.

    Theorem 1.3 (Global stabilization). Suppose the conditions in Theorem 1.1 hold and let (u,v,w) be the solution obtained in Theorem 1.1. If K>α1β1+α2β2 and

    1σ>max0<wK(|d1(w)|2w2β1d1(w)+|d2(w)|2w2β2d2(w))Kw4Kw, (1.9)

    then the co-existence steady state (u,v,w) is globally asymptotically stable.

    The rest of this paper is organized as follows. In Section 2, we present the local existence theorem of solutions and establish some preliminary results. Then in Section 3, we prove Theorem 1.1. We prove Theorem 1.3 in Section 4 and explore the pattern formation in Section 5.

    In the sequel, the integral Ωf(x)dx and fLp(Ω) will be abbreviated as Ωf and fLp, respectively. The generic constants cj or Kj for j=1,2,, are independent of t and will vary in the context. Below we present the local existence result of Eq (1.5), which can be proved in a similar way as [8] by applying Amann's theorem [36,37], and we omit the details for brevity.

    Lemma 2.1 (Local existence). Let ΩR2 be a bounded domain with smooth boundary. Assume that the initial data (u0,v0,w0) satisfies (1.6) and suppose the hypothesis (H0) holds. Then there exists Tmax(0,] such that the system (1.5) admits a unique classical solution (u,v,w)[C(ˉΩ×[0,Tmax))C2,1(ˉΩ×(0,Tmax))]3 satisfying u,v,w>0 for all t>0. Moreover, if Tmax<, then

    limtTmax(u(,t)L+v(,t)L+w(,t)W1,)=.

    Using the similar argument as in [38,Lemma 2.2], we obtain the global boundedness of w immediately.

    Lemma 2.2. Let (u,v,w) be the classical solution of Eq (1.5) obtained in Lemma 2.1. Then it holds that

    0<w(x,t)K:=max{w0L,K}, for allxΩ andt(0,Tmax). (2.1)

    Furthermore, one has

    lim suptw(x,t)K   for all  xˉΩ. (2.2)

    Lemma 2.3. Let (u,v,w) be the classical solution of (1.5) obtained in Lemma 2.1. Assume there is a constant c1>0 such that

    u(,t)Lrc1  for all t(0,Tmax), (2.3)

    then one has

    w(,t)W1,qc2  for all t(0,Tmax), (2.4)

    with

    q{[1,nrnr),if  r<n,[1,),if  r=n,[1,],if  r>n. (2.5)

    Proof. From the third equation of (1.5), we have

    wt=Δww+g(u,v,w)inΩ,wν=0,

    with g(u,v,w):=wβ3uvwu+v+σw(1wK). Since 0<wK (see Eq (2.1)) and u,v>0, we have

    |g(u,v,w)|K+β3Ku+σK+σK2K=K(1+β3u+σ+σKK)K(1+β3+σ+σKK)(u+1),

    which, together with Eq (2.3), gives

    g(u,v,w)Lrc3. (2.6)

    With Eq (2.6), we use the results in [39,Lemma 1] to obtain Eq (2.4) with Eq (2.5) directly.

    Now we will show some basic boundedness properties of the solution (u,v,w) obtained in Lemma 2.1.

    Lemma 2.4. Let (u,v,w) be the classical solution of (1.5) obtained in Lemma 2.1. Then it holds that

    Ωu+ΩvK1  for all t(0,Tmax), (2.7)

    and

    t+τtΩu2+t+τtΩv2K2  for all t(0,Tmaxτ), (2.8)

    where the constants K1,K2>0 are independent of t and τ:=min{1,Tmax/2}.

    Proof. Multiplying the first and second equations of (1.5) by β3, and multiplying the third equation of (1.5) by (β1+β2), then adding the resulting equations, one obtains

    [β3(u+v)+(β1+β2)w]t=Δ[β3(d1(w)u+d2(w)v)+(β1+β2)w]β3(α1u+α2v)+σ(β1+β2)w(1wK). (2.9)

    Thus, integrating Eq (2.9) with respect to x over Ω, we have

    ddtΩ[β3(u+v)+(β1+β2)w]+β3Ω(α1u+α2v)+σ(β1+β2)KΩw2=σ(β1+β2)Ωw. (2.10)

    Using Young's inequality, one has

    (σ+1)(β1+β2)Ωwσ(β1+β2)KΩw2+(σ+1)2(β1+β2)K|Ω|4σ,

    which, substituted into Eq (2.10), gives

    ddtΩ[β3u+β3v+(β1+β2)w]+γ0Ω[β3u+β3v+(β1+β2)w]c1, (2.11)

    where γ0:=min{α1,α2,1} and c1:=(σ+1)2(β1+β2)K|Ω|4σ. Then, applying the Grönwall's inequality to Eq (2.11) gives

    Ω(u+v)Ω(u0+v0)+(β1+β2)β3Ωw0+c1γ0β3,

    which yields Eq (2.7).

    Next, we shall show Eq (2.8) holds based on some ideas in [8]. Under the homogeneous Neumann boundary conditions, we define a shifted Laplacian operator B=Δ+γ1 with γ1:=min{α12d1(0),α22d2(0)}>0. Then B is a sectorial operator in Lp(Ω) for all p(1,) [32], and one can easily show that its inverse B1 satisfies

    B1ϕL2c2ϕL2 for allϕL2(Ω), (2.12)

    and

    B12ϕ2L2c2ϕ2L2 for allϕL2(Ω) (2.13)

    for some constant c2>0. Using the assumptions in (H0) and Eq (2.1), we can derive that there exist two positive constants δ1 and δ2 independent of t such for all j=1,2 one has

    0<δ1|dj(w)|δ2, (2.14)

    and

    0<dj(K)dj(w)dj(0). (2.15)

    Using the definition of B, we can rewrite Eq (2.9) as

    [β3(u+v)+(β1+β2)w]t+B[β3(d1(w)u+d2(w)v)+(β1+β2)w]=(γ1d1(w)α1)β3u+(γ1d2(w)α2)β3v+(γ1+σ)(β1+β2)wσ(β1+β2)Kw2=:f(u,v,w). (2.16)

    Then applying the facts γ1:=min{α12d1(0),α22d2(0)}>0 and Eq (2.15) as well as 0<wK (see Eq (2.1)), we can derive that

    f(u,v,w)(γ1d1(0)α1)β3u+(γ1d2(0)α2)β3v+(γ1+σ)(β1+β2)Kc3,

    where c3:=(γ1+σ)(β1+β2)K>0.

    We multiply Eq (2.16) by B1[β3(u+v)+(β1+β2)w]0 and integrate the result with respect to x over Ω to obtain

    12ddtΩ|B12[β3(u+v)+(β1+β2)w]|2    +Ω[β3(d1(w)u+d2(w)v)+(β1+β2)w][β3(u+v)+(β1+β2)w]    c3ΩB1[β3(u+v)+(β1+β2)w],

    which combined with Eq (2.15) enables us to find a positive constant γ2:=min{d1(K),d2(K),1} such that

    ddtΩ|B12[β3(u+v)+(β1+β2)w]|2+2γ2Ω[β3(u+v)+(β1+β2)w]2    2c3ΩB1[β3(u+v)+(β1+β2)w]. (2.17)

    On one hand, using Eq (2.12), the Hölder inequality and Young's inequality, one has

    2c3ΩB1[β3(u+v)+(β1+β2)w]2c3|Ω|12B1[β3(u+v)+(β1+β2)w]L22c2c3|Ω|12β3(u+v)+(β1+β2)wL2γ22Ω[β3(u+v)+(β1+β2)w]2+2c22c23|Ω|γ2. (2.18)

    On the other hand, using Eq (2.13), we get

    γ22c2Ω|B12[β3(u+v)+(β1+β2)w]|2γ22Ω[β3(u+v)+(β1+β2)w]2. (2.19)

    Defining z(t):=Ω|B12[β3(u+v)+(β1+β2)w]|2, and substituting Eqs (2.18) and (2.19) into Eq (2.17), we obtain

    z(t)+γ22c2z(t)+γ2Ω[β3(u+v)+(β1+β2)w]22c22c23|Ω|γ2. (2.20)

    Then applying the Grönwall's inequality to Eq (2.20) alongside Eq (2.13), one has

    z(t)4c32c23|Ω|γ22+Ω|B12[β3(u0+v0)+(β1+β2)w0]|24c32c23|Ω|γ22+c2Ω|β3(u0+v0)+(β1+β2)w0|2c4. (2.21)

    Integrating Eq (2.20) over (t,t+τ) for τ=min{1,Tmax/2} and applying Eq (2.21), we get

    γ2β23t+τtΩ(u2+v2)γ2t+τtΩ[β3(u+v)+(β1+β2)w]22c22c23|Ω|τγ2+z(t)2c22c23|Ω|γ2+c4,

    which gives Eq (2.8). Then we complete the proof of Lemma 2.4.

    Lemma 2.5. Let (u,v,w) be the classical solution of (1.5) and τ=min{1,Tmax/2}. Then it holds that

    t+τtΩ|w|2K3  for all t(0,Tmaxτ), (2.22)

    and

    wL2K4  for all t(0,Tmax), (2.23)

    as well as

    t+τtΩ|Δw|2K5  for all t(0,,Tmaxτ), (2.24)

    where the positive constants K3, K4 and K5 are independent of t.

    Proof. Multiplying the third equation of (1.5) by w, integrating the resulting equation by parts and using 0<wK in Eq (2.1), we obtain

    ddtΩw2+2Ω|w|2+2β3Ωuvw2u+v+2σKΩw3=2σΩw22σK2|Ω|. (2.25)

    Integrating Eq (2.25) over (t,t+τ) and using the facts 0<τ1 and Eq (2.1), one has Eq (2.22) directly.

    Next, we multiply the third equation of (1.5) by Δw, use the integration by parts to the resulting equation, and apply Young's inequality as well as Eq (2.1) to get

    ddtΩ|w|2+2Ω|Δw|2=2β3Ωuvwu+vΔw+2σKΩw2Δw2σΩwΔwΩ|Δw|2+3β23Ω(uvwu+v)2+3σ2K2Ωw4+3σ2Ωw2Ω|Δw|2+3β23K2Ωu2+3σ2K4|Ω|K2+3σ2K2|Ω|,

    which entails

    ddtΩ|w|2+Ω|Δw|23β23K2Ωu2+c1, (2.26)

    where c1:=3σ2K4|Ω|K2+3σ2K2|Ω|. Applying a Gagliardo-Nirenberg type inequality derived in [16,Lemma 2.5] and noting the fact wL2K|Ω|12, we have

    Ω|w|2c2(ΔwL2wL2+w2L2)c2K|Ω|12(ΔwL2+K|Ω|12)12Ω|Δw|2+(2+c2)c2K2|Ω|2,

    which substituted into Eq (2.26) gives

    ddtΩ|w|2+Ω|w|2+12Ω|Δw|2c3(Ωu2+1), (2.27)

    with c3:=3β23K2+c1+(2+c2)c2K2|Ω|2. For any t(0,Tmax), by Eq (2.22) there exists a t0:=t0(t)((tτ)+,t) such that t00 and

    Ω|w(x,t0)|2dxc4. (2.28)

    Moreover, from Eq (2.8), we know that

    t0+τt0Ωu2(x,s)dxdsK2 for all t0(0,Tmaxτ). (2.29)

    Multiplying Eq (2.27) by et and integrating the results with respect to t over (t0,t), and using the facts Eqs (2.28) and (2.29) and t0<tt0+τt0+1, one has

    w(,t)2L2et0tw(,t0)2L2+etc3tt0es(u(,s)2L2+1)c4+c3tt0(u(,s)2L2+1)c4+c3t0+τt0(u(,s)2L2+1)c4+c3K2+c3.

    Hence Eq (2.23) holds. On the other hand, integrating Eq (2.27) along with Eqs (2.8) and (2.23), it holds that

    t+τtΩ|Δw|22c3t+τtΩu2+2c3τ+2Ω|w(,t)|22c3(K2+1)+2K24,

    which gives Eq (2.24). Then we complete the proof of Lemma 2.5.

    In this section, we shall show the existence of global classical solution based on some ideas in [8]. To this end, we first establish the a priori L2-estimates of u and v.

    Lemma 3.1 (L2-Boundedness). Suppose that the assumptions in Lemma 2.1 hold and let (u,v,w) be the classical solution of (1.5) which is defined on its maximal existence time interval [0,Tmax). Then the solution of (1.5) satisfies

    u(,t)L2+v(,t)L2K6  for all t(0,Tmax), (3.1)

    where the constant K6>0 is independent of t.

    Proof. Multiplying the first equation of (1.5) by u, integrating the results with respect to x over Ω by parts and using the fact u,v>0 for all t(0,Tmax), we have

    12ddtΩu2+α1Ωu2+Ωd1(w)|u|2=β1Ωuvwu+vuΩd1(w)uwuβ1KΩu2+Ωu|d1(w)||w||u|,

    which along with Eqs (2.14) and (2.15) indicates

    12ddtΩu2+α1Ωu2+d1(K)Ω|u|2β1KΩu2+δ2Ωu|w||u|. (3.2)

    Applying Young's inequality and the Hölder inequality, one gets

    δ2Ωu|w||u|d1(K)2Ω|u|2+δ222d1(K)Ωu2|w|2d1(K)2Ω|u|2+δ222d1(K)u2L4w2L4,

    which updates Eq (3.2) as

    ddtu2L2+2α1u2L2+d1(K)u2L22β1Ku2L2+δ22d1(K)u2L4w2L4. (3.3)

    Using the Gagliardo-Nirenberg inequality, the following inequality [16,Lemma 2.5]

    w2L4c1(ΔwL2wL2+w2L2),

    and Eq (2.23), we obtain

    δ22d1(K)u2L4w2L4c2(uL2uL2+u2L2)(ΔwL2K4+K24)c2K4uL2uL2ΔwL2+c2K4u2L2ΔwL2    +c2K24uL2uL2+c2K24u2L2d1(K)u2L2+c3u2L2Δw2L2+c4u2L2, (3.4)

    where c3:=c22K24d1(K) and c4:=c22K442d1(K)+d1(K)2+c2K24.

    Then substituting Eq (3.4) into Eq (3.3), we obtain

    ddtu2L2(2β1K+c4)u2L2+c3u2L2Δw2L2c5u2L2(1+Δw2L2), (3.5)

    where c5:=2β1K+c4+c3. For any t(0,Tmax), by Eq (2.8) there exists a t0:=t0(t)((tτ)+,t) such that t00 and

    Ωu2(x,t0)dxc6. (3.6)

    On the other hand, from Eq (2.24), we know that

    t0+τt0Ω|Δw(x,s)|2dxdsK5, for all t(0,Tmaxτ). (3.7)

    Then integrating Eq (3.5) with respect to t over (t0,t), and using the Eqs (3.6) and (3.7) and tt0+τt0+1, one has

    u(,t)2L2u(,t0)2L2ec5tt0(1+Δw(,s)2L2)dsu(,t0)2L2ec5t0+τt0(1+Δw(,s)2L2)dsc6ec5(1+K5). (3.8)

    Similarly, we can show that

    v(,t)2L2c7. (3.9)

    Then the combination of Eqs (3.8) and (3.9) gives Eq (3.1). The proof of Lemma 3.1 is completed.

    Lemma 3.2 (L-Boundedness). Suppose that the assumptions in Lemma 2.1 hold and let (u,v,w) be the classical solution of (1.5) defined on its maximal existence time interval [0,Tmax). Then it holds that

    u(,t)L+v(,t)LK7  for all t(0,Tmax), (3.10)

    where the constant K7>0 is independent of t.

    Proof. Multiplying the first equation of (1.5) by up1 with p>2 and integrating the resulting equation over Ω by parts, one has

    1pddtΩup+(p1)Ωup2d1(w)|u|2+α1Ωup    =(p1)Ωup1d1(w)uw+β1Ωvwu+vup. (3.11)

    Using Eqs (2.14) and (2.15) and 0<vwu+vwK which is satisfied by the fact u,v>0 for all t(0,Tmax), from Eq (3.11) we get

    1pddtΩup+(p1)d1(K)Ωup2|u|2+α1Ωup    (p1)δ2Ωup1|u||w|+β1KΩup    (p1)d1(K)2Ωup2|u|2+δ22(p1)2d1(K)Ωup|w|2+β1KΩup,

    which together with the fact

    2(p1)d1(K)pΩ|up2|2=p(p1)d1(K)2Ωup2|u|2,

    gives

    ddtΩup+2(p1)d1(K)pΩ|up2|2+α1pΩup    p(p1)δ222d1(K)Ωup|w|2+pβ1KΩup. (3.12)

    Noting the fact uL2K6 in Eq (3.1), and using Lemma 2.3, we can find a constant c1>0 such that

    wL4c1. (3.13)

    Then we can use the Hölder inequality, Gagliardo-Nirenberg inequality and Young's inequality as well as Eq (3.13) to obtain

    p(p1)δ222d1(K)Ωup|w|2p(p1)δ222d1(K)(Ωu2p)12(Ω|w|4)12p(p1)c21δ222d1(K)(Ωu2p)12p(p1)c21δ222d1(K)up22L4c2(up22(11p)L2up22pL4p+up22L4p)c2K6up22(11p)L2+c2Kp6(p1)d1(K)pΩ|up2|2+(c2K6d1(K))pd1(K)p+c2Kp6, (3.14)

    where we have used the fact that

    up2L4p=up2L2Kp26. (3.15)

    Moreover, using the Gagliardo-Nirenberg inequality and Eq (3.15) again, we obtain

           pβ1KΩuppβ1Kup22L2c3(up22(12p)L2up24pL4p+up22L4p)c3K26up22(12p)L2+c3Kp6(p1)d1(K)pΩ|up2|2+(c3K26d1(K))p22d1(K)p+c3Kp6. (3.16)

    Then substituting Eqs (3.14) and (3.16) into Eq (3.12) yields

    ddtΩup+α1pΩupc4(p), (3.17)

    where

    c4(p):=(c2K6d1(K))pd1(K)p+(c3K26d1(K))p22d1(K)p+(c2+c3)Kp6.

    Then applying the Grönwall's inequality to Eq (3.17), we get

    u(,t)pLpu0pLp+c4(p)pα1. (3.18)

    Taking p=4 in Eq (3.18), one enables u(,t)L4c5, which together with Lemma 2.3 gives

    w(,t)Lc6,

    where the constant c6>0 is independent of t and p. Then using the Moser iteration process [33], there exists a constant c7>0 independent of t such that

    u(,t)Lc7. (3.19)

    Similarly, we can find a constant c8>0 such that

    v(,t)Lc8. (3.20)

    Then the combination of Eqs (3.19) and (3.20) gives Eq (3.10). The proof of Lemma 3.2 is completed.

    Proof of Theorem 1.1. From Lemma 3.2, we know there exists a constant c1>0 independent of t such that

    u(,t)L+v(,t)L+w(,t)Lc1,

    which along with the extensibility criterion in Lemma 2.1 and Lemma 2.2 gives Theorem 1.1.

    Under the condition K>α1β1+α2β2, the system (1.5) has only one positive constant steady states (u,v,w) which is defined in Eq (1.8). In what follows, we shall show that the co-existence steady state (u,v,w) is globally asymptotically stable based on the following energy functional:

    E(t):=E(u,v,w)=β2vJu(t)+β1uJv(t)+β1β2(u+v)β3Jw(t), (4.1)

    where

    Jz(t)=Ω(zzzlnzz),   z=u,v,w.

    Then we have the following results.

    Lemma 4.1. Let (u,v,w) be the solution of system (1.5) obtained in Theorem 1.1. If K>α1β1+α2β2 and

    1σ>max0<wK(|d1(w)|2w2β1d1(w)+|d2(w)|2w2β2d2(w))Kw4Kw, (4.2)

    then the co-existence steady states (u,v,w) is globally asymptotically stable.

    Proof. Letting ψ(f)=fflnf and applying Taylor's expansion, for all positive f and f, one has

    ffflnff=ψ(f)ψ(f)=ψ(f)(ff)+12ψ(ξ)(ff)2=f2ξ2(ff)2, (4.3)

    where ξ is between f and f. Then choosing f=u and f=u, we obtain from Eq (4.3) that there exists ξ1 is between u and u such that

    Ju(t)=Ω(uuulnuu)=Ωu2ξ21(uu)20,

    and Ju(t)=0 iff u=u. Using the similar way, one has Jv(t),Jw(t)0 and "=" holds iff v=v,w=w. Hence according to the definition of E(t) in Eq (4.1), we obtain that E(u,v,w)>0 for all (u,v,w)(u,v,w) and moreover E(u,v,w)=0.

    Next, we shall show that ddtE(t)0 under the conditions Eq (4.2) and K>α1β1+α2β2. In fact, using the equations in Eq (1.5), one has

    dE(t)dt=β2vΩ(uuu)ut+β1uΩ(vvv)vt+β1β2(u+v)β3Ω(www)wt=β2vuΩd1(w)|u|2u2β2vuΩd1(w)uwu+β2vΩ(uu)(β1vwu+vα1)    β1uvΩd2(w)|v|2v2β1uvΩd2(w)vwv+β1uΩ(vv)(β2uwu+vα2)    β1β2(u+v)wβ3Ω|w|2w2+β1β2(u+v)β3Ω(ww)(σσwKβ3uvu+v),

    which can be rewritten as

    dE(t)dt=ΩXTBX+β2vΩ(uu)(β1vwu+vα1)+β1uΩ(vv)(β2uwu+vα2)+β1β2(u+v)β3Ω(ww)(σσwKβ3uvu+v)=:ΩXTBX+J1+J2+J3, (4.4)

    where

    X=(uuvvww) and  B=(vud1(w)β20β2vud1(w)w20β1uvd2(w)β1uvd2(w)w2β2vud1(w)w2β1uvd2(w)w2β1β2(u+v)wβ3).

    After some calculations, we can check that the matrix B is positive definite provided

    4(u+v)wβ3uv>|d1(w)|2w2β1d1(w)+|d2(w)|2w2β2d2(w),

    which, together with the fact β3uvu+v=σ(1wK) (see Eq (1.7)), gives

    1σ>Kw4KwF(w), (4.5)

    where the function

    F(w):=|d1(w)|2w2β1d1(w)+|d2(w)|2w2β2d2(w).

    If w0LK, one can easily check that Eq (4.5) holds provided Eq (4.2) is true. Now, we consider the cases w0LK. In fact if Eq (4.2) holds, then there exists a small ε0>0 such that

    1σ>Kw4Kwmax0<wKF(w)+ε0. (4.6)

    From the assumptions (H0), we know that F(w) belongs to C1([0,)), which combined with the fact Eq (2.2), implies that

    lim suptKw4KwF(w)=Kw4Kwlim suptF(w)Kw4Kwmax0<wKF(w),

    and thus for the chosen ε0 in Eq (4.6), there exists a T>0 such that

    Kw4KwF(w)Kw4Kwmax0<wKF(w)+ε0  for all (x,t)ˉΩ×[T,). (4.7)

    Then the combination of Eqs (4.6) and (4.7) gives

    1σ>Kw4KwF(w),  for all (x,t)ˉΩ×[T,).

    Hence the matrix B is positive and there is a constant c1>0 such that

    ΩXTBXc1Ω(|u|2u2+|v|2v2+|w|2w2)for all tT. (4.8)

    On the other hand, using the fact α1β1=vwu+v (see Eq (1.7)), one gets

    J1:=β2vΩ(uu)(β1vwu+vα1)=β1β2vΩ(uu)(vwu+vα1β1)=β1β2vΩ(uu)(vwu+vvwu+v)=β1β2vu+vΩ(uu)vwu+vwvuvwvvwu+v,

    which together with the fact

    vwu+vwvuvwvvw=vv(ww)+vwuuvw=vv(ww)+vu(ww)+vwuuvw=v(u+v)(ww)+vwuvwu+vwuuvw=v(u+v)(ww)vw(uu)+uw(vv),

    gives

    J1=β1β2vΩv(uu)(ww)u+v+β1β2vwu+vΩu(uu)(vv)u+vβ1β2vwu+vΩv(uu)2u+v. (4.9)

    Similarly, we can use the fact α2β2=uwu+v (see Eq (1.7)) to obtain

    J2:=β1uΩ(vv)(β2uwu+vα2)=β1β2uΩu(vv)(ww)u+v+β1β2uwu+vΩv(uu)(vv)u+v    β1β2uwu+vΩu(vv)2u+v. (4.10)

    At last, using the fact σ=β3uvu+v+σwK (see Eq (1.7)), we have

    J3:=β1β2(u+v)β3Ω(ww)(σσwKβ3uvu+v)=β1β2(u+v)β3Ω(ww)(β3uvu+vβ3uvu+v)β1β2(u+v)σβ3KΩ(ww)2=β1β2(u+v)Ω(ww)uvu+uvvuvuuvv(u+v)(u+v)    β1β2(u+v)σβ3KΩ(ww)2=β1β2uΩu(vv)(ww)u+vβ1β2vΩv(uu)(ww)u+v    β1β2(u+v)σβ3KΩ(ww)2. (4.11)

    Combining Eqs (4.9) and (4.10) with Eq (4.11), we have

    3j=1Jj=β1β2wu+vΩ(vu+uv)(uu)(vv)u+v    β1β2wu+vΩvv(uu)2+uu(vv)2u+vβ1β2(u+v)σβ3KΩ(ww)2=β1β2wu+vΩf(u,v)u+vβ1β2(u+v)σβ3KΩ(ww)2, (4.12)

    where

    f(u,v)=(vu+uv)(uu)(vv)vv(uu)2uu(vv)2=v(uu)(uu)(vv)+vu(uu)(vv)    +u(vv)(uu)(vv)+uv(uu)(vv)    vv(uu)2uu(vv)2=v(uu)2(vv)+2vu(uu)(vv)    +u(uu)(vv)2vv(uu)2uu(vv)2=v2(uu)2u2(vv)2+2uv(uu)(vv)=(vuuv)2.

    Then we can update Eq (4.12) as

    3j=1Jj=β1β2wu+vΩ(vuuv)2u+vβ1β2(u+v)σβ3KΩ(ww)2. (4.13)

    Substituting Eqs (4.8) and (4.13) into Eq (4.4) gives for all tT

    dE(t)dtβ1β2wu+vΩ(vuuv)2u+vβ1β2(u+v)σβ3KΩ(ww)2    c1Ω(|u|2u2+|v|2v2+|w|2w),

    which indicates dE(t)dt0 for all u,v,w. Moreover, one can check that if dE(t)dt=0, then u=v=0 and w=w, which implies that

    u=˜u,  v=˜v

    with ˜u and ˜v are some positive constants. Since (u,v,w)=(˜u,˜v,w) is a solution of (1.5), then one has

    {β1˜vw˜u+˜vα1=0,β2˜uw˜u+˜vα2=0,β3˜u˜v˜u+˜vσ+σwK=0,

    which gives

    ˜u=σβ3(1wK)(1+α2β1α1β2)=u,  ˜v=σβ3(1wK)(1+α1β2α2β1)=v,

    Thus dE(t)dt=0 if and only if (u,v,w)=(u,v,w). By using the LaSalle's invariant principle (cf. [41,Theorem 5.24]), we get that the co-existence steady state (u,v,w) is globally asymptotically stable.

    Proof of Theorem 1.3. Theorem 1.3 is a consequence of Lemma 4.1.

    In this section, we assume K>α1β1+α2β2 and will study the effect of density-suppressed dispersals on the dynamics of the system (1.5).

    Denote dj:=dj(w),dj:=dj(w)(j=1,2) and linearize (1.5) at (u,v,w) to obtain

    {Φt=AΔΦ+BΦ,xΩ,t>0,(ν)Φ=0,xΩ,t>0,Φ(x,0)=(u0u,v0v,w0w)T,xΩ, (5.1)

    where T denotes the transpose and

    Φ=(uuvvww),A=(d10d1u0d2d2v001),  B=(Bij)3×3

    with

    B11=α1uu+v<0,B12=α2β1uβ2(u+v)>0,B13=β1uvu+v>0,B21=α1β2vβ1(u+v)>0,B22=α2vu+v<0,B23=β2uvu+v>0,B31=α1β3vβ1(u+v)<0,B32=α2β3uβ2(u+v)<0,B33=σwK<0.

    Then by the standard linear stability principle, the linear stability of (u,v,w) is determined by the eigenvalues of the matrix (μkA+B) where the sequence {μk}k=0 with 0=μ0<μ1μ2μ3 denotes the eigenvalues of Δ under the Neumann boundary condition. The characteristic equation of (μkA+B) is

    ρ3+D1(μk)ρ2+D2(μk,|dj|)ρ+D3(μk,|dj|)=0, (5.2)

    where

    D1(μk):=μk(d1+d2+1)+σwK+α1u+α2vu+v>0,
    D2(μk,|dj|):=μ2k(d1d2+d1+d2)+μk{(d1+d2)σwK+(d1+1)α2vu+v+(d2+1)α1uu+v}+μk{β3uvu+v(α1d1β1α2d2β2)}+β3uv(α1v+α2u)(u+v)2+σw(α1u+α2v)K(u+v)>0,
    D3(μk,|dj|):=μ3kd1d2+μ2k{d1d2σwK+d1α2vu+v+d2α1uu+v+β3uvu+v(α1d1d2β1α2d1d2β2)}+μk{d1α2vu+v(σwK+β3u2u+v)+d2α1uu+v(σwK+β3v2u+v)}+μk{α1α2β3uvu+v(d1β1d2β2)}+α1α2β3uvu+v>0.

    Hence, by direct calculations, one has

    H(μk,|dj|):=D1(μk)D2(μk,|dj|)D3(μk,|dj|)=γ1μ3k+γ2μ2k+γ3μk+A1A2A3+A0(μk,|dj|)+A4(μk,|dj|), (5.3)

    where γ1,γ2,γ3 are positive constants and A1A2A3>0 are independent of dj(j=1,2) and μk (see details in the Appendix for the precise definitions of γi and Ai, i=1,2,3), and

    A0(μk,|dj|):=μ2kβ3uvu+v{(d1+1)α1d1β1(d2+1)α2d2β2}0,

    as well as

    A4(μk,|dj|):=μkβ3uv(u+v){(d1)α1β1(σwK+(α1α2)uu+v)+(d2)α2β2(σwK+(α2α1)vu+v)}.

    Then using the Routh-Hurwitz criterion, we have the results stated as below.

    Proposition 5.1. The (u,v,w) is linearly stable if one of the following conditions holds:

    1) d1(w),d2(w) are constants with σ>0;

    2) dj(w)<0 with σ>0 if α1=α2 or σKβ1β2f(β)|α1α2|(α1β2+α2β1)2 if α1α2, where

    f(β):={α2β1,ifα1<α2,α1β2,ifα1>α2, (5.4)

    and j=1,2.

    Proof. In the case that dj(w)(j=1,2) are constants, one has

    A0(μk,|dj|)=A4(μk,|dj|)=0.

    On the other hand, considering the case dj(w)<0. If α1=α2, we get

    A0(μk,|dj|),A4(μk,|dj|)>0.

    As for α1α2 with

    σKβ1β2f(β)|α1α2|(α1β2+α2β1)2,

    we can calculate to get

    A0(μk,|dj|)>0,A4(μk,|dj|)0.

    All cases discussed above indicate D1(μk)D2(μk,|dj|)D3(μk,|dj|)>0 for all kN. Then by the Routh-Hurwitz criterion [34], we know (u,v,w) is linearly stable.

    Hence, we are left to study the possible patterns bifurcating from the co-existence steady state (u,v,w) in the following range of parameters:

    (H1) α1α2, 0<σ<Kβ1β2f(β)|α1α2|(α1β2+α2β1)2 and di(w)<0, where i=1 if α1<α2 or i=2 if α1>α2 and f(β) is defined in (5.4).

    Proposition 5.2. Let σ,K,α1,α2,β1,β2 and β3 be fixed positive parameters, and suppose the assumption (H1) holds. Then (u,v,w) is linearly unstable provided that |d1(w)| is large enough if α1<α2 (or |d2(w)| is large enough if α2<α1) and there is some k01 such that

    0<μk0<{11+d1|σwK+(α1α2)uu+v|,ifα1<α2,11+d2|σwK+(α2α1)vu+v|,ifα1>α2. (5.5)

    Proof. Noting that H(μk,|dj|) can be rewritten as

    H(μk,|dj|)=γ1μ3k+γ2μ2k+γ3μk+A1A2A3  +(d1)α1β3uvβ1(u+v){(d1+1)μ2k+(σwK+(α1α2)uu+v)μk}  +(d2)α2β3uvβ2(u+v){(d2+1)μ2k+(σwK+(α2α1)vu+v)μk}.

    Since (H1) holds, then for some fixed μk0>0 satisfying Eq (5.5), we have

    H(μk,|dj|):=D1(μk)D2(μk,|dj|)D3(μk,|dj|)<0

    when |d1(w)| is large enough if α1<α2 (or |d2(w)| is large enough if α2<α1) and (u,v,w) is linearly unstable by the Routh-Hurwitz criterion [40].

    Remark 5.3. The above results imply that the density-suppressed diffusion can induce the instability of (u,v,w) and trigger the pattern formation.

    In this subsection, we aim to give some numerical simulations to complement the previous analysis in Section 5.1. Without loss of generality, we assume α1<α2, and define the set

    H={(|d1|,η)R2+:H(|d1|,η)=0}

    as the bifurcation curve [35], where the linearized system (5.1) admits an eigenvalue with zero real part and the curve H is the graph of the function

    dH(η)=β1(u+v)α1β3uv((d1+1)ησwK(α1α2)uu+v)1{(η2γ1+ηγ2+γ3+A1A2A3η)+(d2)α2β3uvβ2(u+v)((d2+1)η+σwK+(α2α1)vu+v)}, (5.6)

    where the constants γj,Aj>0(j=1,2,3) are given in the Appendix. Then the characteristic Eq (4.3) has a pair of purely imaginary eigenvalues if |d1(w)|=dH(μk0) for some k0N and μk0 satisfying Eq (5.5) and the Hopf bifurcation may emerge.

    We first fix parameter values as follows:

    β1=0.3,β2=1,β3=1.3,α1=0.1,α2=1,K=2. (5.7)

    Then by the assumption (H1), one has

    0<σ<8180.

    Hence, we fix σ=110 in what follows without loss of generality. By Eq (1.8) we have

    u=439, v=4117, w=43. (5.8)

    In addition, we choose the motility function d1(w) as

    d1(w)=eD(w43),

    where the constant D>0 and d1(w)=DeD(w43). Then Eq (5.5) is equivalent to

    0<μk0<73240.

    Using Theorem 1.3 and Proposition 5.2, the co-existence steady state (439,4117,43) may lose its stability when the parameter D>0 satisfies

    10D23max0<wKeD(w43)w2+max0<wK|d2(w)|2w2d2(w)160, (5.9)

    and

    |d1(43)|=DdH(η) (5.10)

    with 0<η<73240.

    In our numerical simulation, we set the initial value (u0,v0,w0) as a small random perturbation of (439,4117,43) and hence K=2. Moreover, we perform numerical simulations in one dimension and set Ω=(0,16π). Choosing D>0 satisfies (5.9) and (5.10) with η=(k/16)2<73240, which indicates allowable unstable modes are k=1,2,3,4,5,6,7,8. Next, we will consider the following two cases for d2(w).

    Case 1: d2(w)=1. In this case, d2=d2(w)=0 and according to the definition of dH(η) in Eq (5.6), we can calculate to obtain that

    dH(η)=90(731202η)1(8η2+4715η+28937200+73348000η),

    and Eq (5.9) can be updated as

    e43D12e2

    with D>4. Then after some calculations, one gets dH(1256)647.963, dH(4256)223.026, dH(9256)159.956, dH(16256)162.600, dH(25256)204.933, dH(36256)305.214, dH(49256)548.498 and dH(64256)1450.71. Hence, dH(9256)159.956 is the critical value for possible pattern formation. In our numerical simulations shown in Figure 1, we choose D=180 and find the spatio-temporal pattern. In particular the time evolutionary profiles of solutions are oscillatory, which implies the bifurcation might be Hopf bifurcation.

    Figure 1.  Numerical simulation of spatio-temporal patterns (first panel) and time evolutionary profiles (second panel) generated by the model (1.5), where d1(w)=eD(w43) with D=180 and d2(w)=1, other parameter values are given in Eq (5.7) with σ=0.1, and the initial value (u0,v0,w0) is chosen as a small random perturbation of (u,v,w) given in Eq (5.8).

    Case 2: d2(w)=e50(w43). For this case, d2=d2(w)=50 and hence

    dH(η)=90(731202η)1(8η2+9715η+21312400+73348000η),

    and Eq (5.9) can be simplified as

    e2003+103e43D40e2.

    One also can calculate to get that dH(1256)722.767, dH(4256)306.961, dH(9256)260.876, dH(16256)291.910, dH(25256)381.793, dH(36256)567.953, dH(49256)997.112 and dH(64256)2546.860. Therefore, dH(9256)260.876 is the critical value for possible pattern formation. In our numerical simulations, we choose D=268 and find the spatio-temporal pattern shown in Figure 2. Again oscillatory time evolutionary profiles indicate the solution bifurcating from the co-existence steady state may undergo a Hopf bifurcation.

    Figure 2.  Numerical simulation of spatio-temporal patterns (first panel) and oscillatory time-evolutionary profiles (second panel) generated by the model (1.5), where d1(w)=eD(w43) with D=268 and d2(w)=e50(w43), other parameter values are given in Eq (5.7) with σ=0.1, and the initial value is chosen as a small random perturbation of (u,v,w) given in Eq (5.8).

    The authors are grateful to four referees for their valuable comments, which greatly improved the exposition of our paper. The research of H. Y. Jin was supported by the NSF of China (No. 11871226), Guangdong Basic and Applied Basic Research Foundation (No. 2020A1515010140, No. 2022B1515020032), Guangzhou Science and Technology Program (No. 202002030363).

    The authors declare there is no conflict of interest.

    Below we show the details for the parameters γ1,γ2,γ3>0 and A1A2A3>0 present in Eq (5.3). By direct calculations, we obtain

    γ1=(d1+d2+1)(d1d2+d1+d2)d1d2=(d1+d2)(d1d2+d1+d2)+(d1+d2)=(d1d2+d1+d2+1)(d1+d2)>0.

    As for

    γ2=(d1+d2+1){(d1+d2)σwK+(d1+1)α2vu+v+(d2+1)α1uu+v}    +(d1d2+d1+d2)(σwK+α1u+α2vu+v)(d1d2σwK+d1α2vu+v+d2α1uu+v)=(d1+d2+1){(d1+d2)σwK+(d1+1)α2vu+v+(d2+1)α1uu+v}    +d1d2(α1u+α2v)u+v+(d1+d2)σwK+d1α1u+d2α2vu+v,

    which implies γ2>0 and is independent of dj(j=1,2) and μk. Now, we calculate the other parameters:

    γ3=(σwK+α1u+α2vu+v){(d1+d2)σwK+(d1+1)α2vu+v+(d2+1)α1uu+v}+(d1+d2+1){β3uv(α1v+α2u)(u+v)2+σw(α1u+α2v)K(u+v)}d1α2vu+v(σwK+β3u2u+v)d2α1uu+v(σwK+β3v2u+v)=d2α2β3u2v(u+v)2+d1α1β3uv2(u+v)2+β3uv(α1v+α2u)(u+v)2+2(d1+d2+1)σw(α1u+α2v)K(u+v)+(d1+d2)(σwK)2+((d1+1)α2v+(d2+1)α1u)α1u+α2v(u+v)2>0,

    and

    A1:=α1u+α2vu+v+σwK>0,A2:=β3uv(α1v+α2u)(u+v)2+σw(α1u+α2v)K(u+v)>0,A3:=α1α2β3uvu+v>0.

    Hence, a direct calculation gives

    A1A2A3>(α1u+α2v)(α1v+α2u)β3uv(u+v)3α1α2β3(uv3+u3v)+2α1α2β3u2v2(u+v)3=(α1α2)2β3u2v2(u+v)30.


    [1] M. Saleem, A. K. Tripathi, A. H. Sadiyal, Coexistence of species in a defensive switching model, Math. Biosci., 181 (2003), 145–164. https://doi.org/10.1016/S0025-5564(02)00152-9 doi: 10.1016/S0025-5564(02)00152-9
    [2] S. Takahashi, M. Hori, Unstable evolutionarily stable strategy and oscillation: A model of lateral asymmetry in scale-eating cichlids, Am. Nat., 144 (1994), 1001–1020. https://doi.org/10.1086/285722 doi: 10.1086/285722
    [3] P. Y. H. Pang, M. Wang, Strategy and stationary pattern in a three-species predator-prey model, J. Differ. Equations, 200 (2004), 245–273. https://doi.org/10.1016/j.jde.2004.01.004 doi: 10.1016/j.jde.2004.01.004
    [4] Y. Cai, Q. Cao, Z. A. Wang, Asymptotic dynamics and spatial patterns of a ratio-dependent predator-prey system with prey-taxis, Appl. Anal., 101 (2022), 81–99. https://doi.org/10.1080/00036811.2020.1728259 doi: 10.1080/00036811.2020.1728259
    [5] J. Wang, X. Guo, Dynamics and pattern formations in a three-species predator-prey model with two prey-taxis, J. Math. Anal. Appl., 475 (2019), 1054–1072. https://doi.org/10.1016/j.jmaa.2019.02.071 doi: 10.1016/j.jmaa.2019.02.071
    [6] X. Guo, J. Wang, Dynamics and pattern formations in diffusive predator-prey models with two prey-taxis, Math. Methods Appl. Sci., 42 (2019), 4197–4212. https://doi.org/10.1002/mma.5639 doi: 10.1002/mma.5639
    [7] P. Kareiva, G. Odell, Swarms of predators exhibit "preytaxis" if individual predators use area-restricted search, Am. Nat., 130 (2015), 233–270. https://doi.org/10.1086/284707 doi: 10.1086/284707
    [8] H. Y. Jin, Z. A. Wang, Global dynamics and spatio-temporal patterns of predator-prey systems with density-dependent motion, European J. Appl. Math., 32 (2021), 652–682. https://doi.org/10.1017/S0956792520000248 doi: 10.1017/S0956792520000248
    [9] Z. A. Wang, J. Xu, On the Lotka-Volterra competition system with dynamical resources and density-dependent diffusion, J. Math. Biol., 82 (2021), 1–37. https://doi.org/10.1007/s00285-021-01562-w doi: 10.1007/s00285-021-01562-w
    [10] E. Keller, L. Segel, Traveling bands of chemotactic bacteria: A theoretical analysis, J. Theor. Biol., 30 (1971), 377–380. https://doi.org/10.1016/0022-5193(71)90051-8 doi: 10.1016/0022-5193(71)90051-8
    [11] E. Keller, L. Segel, Model for chemotaxis, J. Theor. Biol., 30 (1971), 225–234. https://doi.org/10.1016/0022-5193(71)90050-6
    [12] X. Fu, L. H. Tang, C. Liu, J. D. Huang, T. Hwa, P. Lenz, Stripe formation in bacterial systems with density-suppressed motility, Phys. Rev. Lett., 108 (2012), 198102. https://doi.org/10.1103/physrevlett.108.198102 doi: 10.1103/physrevlett.108.198102
    [13] C. Liu, X. Fu, L. Liu, X. Ren, C. K. Chau, S. Li, et al., Sequential establishment of stripe patterns in an expanding cell population, Science, 334 (2011), 238–241. https://doi.org/10.1126/science.1209042 doi: 10.1126/science.1209042
    [14] J. Ahn, C. Yoon, Global well-posedness and stability of constant equilibria in parabolic-elliptic chemotaxis systems without gradient sensing, Nonlinearity, 32 (2019), 1327–1251. https://doi.org/10.1088/1361-6544/aaf513 doi: 10.1088/1361-6544/aaf513
    [15] M. Burger, P. Laurençot, A. Trescases, Delayed blow-up for chemotaxis models with local sensing, J. Lond. Math. Soc., 103 (2021), 1596–1617. https://doi.org/10.1112/jlms.12420 doi: 10.1112/jlms.12420
    [16] H. Y. Jin, Y. J. Kim, Z. A. Wang, Boundedness, stabilization, and pattern formation driven by density-suppressed motility, SIAM J. Appl. Math., 78 (2018), 1632–1657. https://doi.org/10.1137/17M1144647 doi: 10.1137/17M1144647
    [17] K. Fujie, J. Jiang, Global existence for a kinetic model of pattern formation with density-suppressed motilities, J. Differ. Equations, 269 (2020), 5338–5378. https://doi.org/10.1016/j.jde.2020.04.001 doi: 10.1016/j.jde.2020.04.001
    [18] J. Jiang, P. Laurençot, Y. Zhang, Global existence, uniform boundedness, and stabilization in a chemotaxis system with density-suppressed motility and nutrient consumption, Comm. Partial Differ. Equations, 47 (2022), 1024–1069. https://doi.org/10.1080/03605302.2021.2021422 doi: 10.1080/03605302.2021.2021422
    [19] J. Jiang, P. Laurençot, Global existence and uniform boundedness in a chemotaxis model with signal-dependent motility, J. Differ. Equations, 299 (2021), 513–541. https://doi.org/10.1016/j.jde.2021.07.029 doi: 10.1016/j.jde.2021.07.029
    [20] H. Y. Jin, Z. A. Wang, Critical mass on the Keller-Segel system with signal-dependent motility. Proc. Amer. Math. Soc., 148 (2020), 4855–4873. https://doi.org/10.1090/proc/15124
    [21] H. Y. Jin, S. Shi, Z. A. Wang, Boundedness and asymptotics of a reaction-diffusion system with density-dependent motility, J. Differ. Equations, 269 (2020), 6758–6793. https://doi.org/10.1016/j.jde.2020.05.018 doi: 10.1016/j.jde.2020.05.018
    [22] W. Lv, Q. Wang, An n-dimensional chemotaxis system with signal-dependent motility and generalized logistic source: global existence and asymptotic stabilization, Proc. Roy. Soc. Edinburgh Sect. A, 151 (2021), 821–841. https://doi.org/10.1017/prm.2020.38 doi: 10.1017/prm.2020.38
    [23] W. Lyu, Z.-A. Wang, Global classical solutions for a class of reaction-diffusion system with density-suppressed motility, Electron. Res. Arch., 30 (2022), 995–1015. https://doi.org/10.3934/era.2022052 doi: 10.3934/era.2022052
    [24] W. Lv, Global existence for a class of chemotaxis-consumption systems with signal-dependent motility and generalized logistic source, Nonlinear Anal. Real World Appl., 56 (2020), 103160. https://doi.org/10.1016/j.nonrwa.2020.103160 doi: 10.1016/j.nonrwa.2020.103160
    [25] J. Smith Roberge, D. Iron, T. Kolokolnikov, Pattern formation in bacterial colonies with density-dependent diffusion, European J. Appl. Math., 30 (2019), 196–218. https://doi.org/10.1017/S0956792518000013 doi: 10.1017/S0956792518000013
    [26] M. Wang, J. Wang, Boundedness in the higher-dimensional Keller-Segel model with signal-dependent motility and logistic growth, J. Math. Phys., 60 (2019), 011507. https://doi.org/10.1063/1.5061738 doi: 10.1063/1.5061738
    [27] C. Yoon, Y.-J. Kim, Global existence and aggregation in a Keller-Segel model with Fokker-Planck diffusion, Acta Appl. Math., 149 (2017), 101–123. https://doi.org/10.1007/s10440-016-0089-7 doi: 10.1007/s10440-016-0089-7
    [28] C. Xu, Y. Wang, Asymptotic behavior of a quasilinear Keller-Segel system with signal-suppressed motility, Calc. Var. Partial Differ. Equations, 60 (2021), 1–29. https://doi.org/10.1007/s00526-021-02053-y doi: 10.1007/s00526-021-02053-y
    [29] J. Li, Z.-A. Wang, Traveling wave solutions to the density-suppressed motility model, J. Differential Equations, 301 (2021), 1–36. https://doi.org/10.1016/j.jde.2021.07.038 doi: 10.1016/j.jde.2021.07.038
    [30] M. Ma, R. Peng, Z. Wang. Stationary and non-stationary patterns of the density-suppressed motility model. Phys. D, 402 (2020), 132259. https://doi.org/10.1016/j.physd.2019.132259
    [31] Z.-A. Wang, X. Xu, Steady states and pattern formation of the density-suppressed motility model, IMA J. Appl. Math., 86 (2021), 577–603. https://doi.org/10.1093/imamat/hxab006 doi: 10.1093/imamat/hxab006
    [32] A. Yagi, Abstract parabolic evolution equations and their applications, Springer Science and Business Media, 2009. https://doi.org/10.1007/978-3-642-04631-5
    [33] N. Alikakos, Lp bounds of solutions of reaction-diffusion equations, , Commun. Partial Differ. Equations, 4 (1979), 827–868. https://doi.org/10.1080/03605307908820113 doi: 10.1080/03605307908820113
    [34] J. Murray, Mathematical Biology I: An Introduction, 3rd edition, Springer, Berlin, 2002. https://doi.org/10.1007/b98868
    [35] J. Wang, J. Shi, J. Wei, Dynamics and pattern formation in a diffusive predator-prey system with strong Allee effect in prey, J. Differ. Equations, 251 (2011), 1276–1304. https://doi.org/10.1016/j.jde.2011.03.004 doi: 10.1016/j.jde.2011.03.004
    [36] H. Amann, Dynamic theory of quasilinear parabolic equations. II. Reaction-diffusion systems, Differ. Integral Equations, 3 (1990), 13–75.
    [37] H. Amann, Nonhomogeneous linear and quasilinear elliptic and parabolic boundary value problems, Funct. Spaces Differ. Oper. Nonlinear Anal., 133 (1993), 9–126. https://doi.org/10.1007/978-3-663-11336-2_1 doi: 10.1007/978-3-663-11336-2_1
    [38] H. Y. Jin, Z. A. Wang, Global stability of prey-taxis systems, J. Differ. Equations, 262 (2017), 1257–1290. https://doi.org/10.1016/j.jde.2016.10.010 doi: 10.1016/j.jde.2016.10.010
    [39] R. Kowalczyk, Z. Szymańska, On the global existence of solutions to an aggregation model, J. Math. Anal. Appl., 343 (2008), 379–398. https://doi.org/10.1016/j.jmaa.2008.01.005 doi: 10.1016/j.jmaa.2008.01.005
    [40] P. Liu, J. Shi, Z.-A. Wang, Pattern formation of the attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2597–2625. https://doi.org/10.3934/dcdsb.2013.18.2597 doi: 10.3934/dcdsb.2013.18.2597
    [41] S. Sastry, Nonlinear System: Analysis, Stability, and Control, Springer, New York, 1999. https://doi.org/10.1007/978-1-4757-3108-8
  • This article has been cited by:

    1. Zhoumeng Xie, Yuxiang Li, Global solutions near homogeneous steady states in a fully cross-diffusive predator–prey system with density-dependent motion, 2023, 74, 0044-2275, 10.1007/s00033-023-02127-1
  • Reader Comments
  • © 2022 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(2393) PDF downloads(182) Cited by(1)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog