
In this paper, we concern with the predator-prey system with generalist predator and density-dependent prey-taxis in two-dimensional bounded domains. We derive the existence of classical solutions with uniform-in-time bound and global stability for steady states under suitable conditions through the Lyapunov functionals. In addition, by linear instability analysis and numerical simulations, we conclude that the prey density-dependent motility function can trigger the periodic pattern formation when it is monotone increasing.
Citation: Tingfu Feng, Leyun Wu. Global dynamics and pattern formation for predator-prey system with density-dependent motion[J]. Mathematical Biosciences and Engineering, 2023, 20(2): 2296-2320. doi: 10.3934/mbe.2023108
[1] | 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 |
[2] | 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 |
[3] | 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 |
[4] | Nazanin Zaker, Christina A. Cobbold, Frithjof Lutscher . The effect of landscape fragmentation on Turing-pattern formation. Mathematical Biosciences and Engineering, 2022, 19(3): 2506-2537. doi: 10.3934/mbe.2022116 |
[5] | 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 |
[6] | 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 |
[7] | Xiaomei Bao, Canrong Tian . Turing patterns in a networked vegetation model. Mathematical Biosciences and Engineering, 2024, 21(11): 7601-7620. doi: 10.3934/mbe.2024334 |
[8] | 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 |
[9] | 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 |
[10] | Yue Xing, Weihua Jiang, Xun Cao . Multi-stable and spatiotemporal staggered patterns in a predator-prey model with predator-taxis and delay. Mathematical Biosciences and Engineering, 2023, 20(10): 18413-18444. doi: 10.3934/mbe.2023818 |
In this paper, we concern with the predator-prey system with generalist predator and density-dependent prey-taxis in two-dimensional bounded domains. We derive the existence of classical solutions with uniform-in-time bound and global stability for steady states under suitable conditions through the Lyapunov functionals. In addition, by linear instability analysis and numerical simulations, we conclude that the prey density-dependent motility function can trigger the periodic pattern formation when it is monotone increasing.
The dynamical relationship between predators and their prey is one of the dominant themes in ecology. The origin and theory of predator-prey model is due to pioneer work of Lotka and Volterra [1,2]. There have been a lot of studies on the dynamics of this particular type of model through developing various modifications of mathematical models of prey-predator interactions (e.g., [3,4,5,6,7,8]).
The non-random foraging strategies in the predator-prey dynamics, prey-taxis allows predators to move towards regions of higher prey density and to search more actively for prey. Such a prey-taxis model was derived by Kareiva and Odell in [9], and they studied predator aggregation in high prey density areas. It can be described as:
{ut=∇⋅(d(v)∇u)−∇⋅(uχ(v)∇v)+H1(u,v),vt=DΔv+H2(u,v), | (1.1) |
where u(x,t) and v(x,t) denote the population density of predators and preys at position x and time t respectively, and D is a positive constant standing for the diffusion rate of the prey. The terms ∇⋅(d(v)∇u) and −∇⋅(uχ(v)∇v) account for the diffusion of predators with coefficient d(v) and the prey-taxis with coefficient χ(v) respectively. H1(u,v) and H2(u,v) denote the predator-prey interactions. There mainly are three kinds of typical interspecific interactions: predator-prey, competition and mutualism, which can be represented as
H1(u,v)=f(u)+c1uF(v),H2(u,v)=g(v)−c2uF(v), |
where the functions f(u) and g(v) stand for the intra-specific interactions of predators and prey respectively. The parameters c1 and c2 are positive constants representing the coefficients of inter-specific interactions of predators and prey, and F(v) is the so-called functional response function.
In particular, if χ(v)=−d′(v), the system (1.1) can be written as
{ut=Δ(d(v)u)+H1(u,v),vt=DΔv+H2(u,v), | (1.2) |
the diffusion term Δ(d(v)u) with d′(v)<0 is called the "density-suppressed motility" (see [10,11,12,13,14,15]), which can also characterize the incessant tumbling of cells at high concentration, resulting in a vanishing macroscopic motility. Here d(v) is called the motility function, d′(v)<0 means that the predator reduce its motility when encountering the prey. When H1(u,v)=0, H2(u,v)=u−v and d(v)=c0v−k decays algebraically in v, the solution may exist globally in two or higher dimensions. For example, Yoon and Kim in [16] proved that system (1.2) has a unique global bounded classical solution for any k>0 under a smallness assumption on c0 in any dimensions. The only global existence result without smallness assumptions was recently given by Ahn and Yoon [17]. Under the assumptions that there exist positive constants γ1,γ2,γ3 such that γ1≤d(v)≤γ2 and |d′(v)|≤γ3, Tao and Winkler in [18] proved the existence of global classical solutions in the 2-dimensional case and global weak solutions in 3-dimensions. While if d(v) decays exponentially, the solution may blow-up in two dimensions with a critical mass, see [19,20,21] and so on. For H1(u,v)≠0 with logistic growth on f(u)=μu(1−u), there also are many interesting results. The global existence and asymptotic behavior of solutions was first established by Jin, Kim and Wang in [12] under certain conditions on d(v) in two dimensions, which has been developed by many authors, please refer to [19,22,23,24,25,26,27,28] and so on.
In the above mentioned references, the authors assumed d′(v)<0. While under normal circumstances, the predators will increase their motilities and keep chasing (such as wolves and sheep) when encountering the prey until they succeed. Therefore, it is meaningful for us to consider the case d′(v)≥0, and we shall consider general case for d(v) without monotonicity assumptions in this paper.
In this paper, we consider the following density-dependent predator-prey system:
{ut=Δ(d(v)u)+u(a1−b1u)+αuF(v),x∈Ω,t>0,vt=Δv+v(a2−b2v)−uF(v),x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t>0,(u,v)(x,0)=(u0,v0)(x),x∈Ω, | (1.3) |
where a1,a2>0 represent the intrinsic growth rates of species, b1,b2>0 are the death rates due to intra-specific competition and α>0 denotes the intrinsic predation rate. F(v) is the so-called functional response function accounting for the intake rate of predators as a function of prey density, and d(v) is the motility function as we mentioned above. The most common types F(v) in the literature are
● F(v)=v (Lotka-Volterra type or Holling type Ⅰ);
● F(v)=vλ+v (Holling type Ⅱ);
● F(v)=vmλm+vm (Holling type Ⅲ) with constants λ>0 and m>1.
We make the following assumptions throughout the whole paper:
(H1) d(v)∈C3([0,∞)) and d(v)>0 on [0,∞).
(H2) F(v)∈C1([0,∞)),F(0)=0,F(v)>0 in (0,∞) and F′(v)>0.
In (1.3), the predator is called the specialist predator if a1<0, Jin and Wang investigate the global boundedness, asymptotic stability and pattern formation of system (1.3) [29]. They study the dynamic behaviors of the predator and the prey under the condition d′(v)<0. In this paper, we assume a1>0 (the corresponding predator is called the generalist predator) and make no assumptions on the monotonicity of d(v). Usually, a generalist species is able to thrive in a wide variety of environmental conditions and can make use of a variety of different resources (for example, a heterotroph with a varied diet). A specialist species can thrive only in a narrow range of environmental conditions or has a limited diet. Most organisms do not all fit neatly into either group. Some species are highly specialized (the most extreme case being monophagous, eating one specific type of food), others less so, and some can tolerate many different environments. In other words, there is a continuum from highly specialized to broadly generalist species. We mainly focus on exploring the global dynamics and spatial-temporal patterns for generalist predators with density-dependent motion for more general motility functions d(v). We mention here Nakashima and Yamada in [30] studied the existence of positive solutions for boundary value problems of nonlinear elliptic systems which arise in the study of the Lotka-Volterra prey-predator models with cross-diffusion in the special case d(v)=1+αv.
We first derive the global boundedness and existence results for the classical solutions to the system (1.3).
Theorem 1.1 (Global boundedness). Let Ω⊂R2 be a bounded domain with smooth boundary and the hypotheses (H1)-(H2) hold. Assume (u0,v0)∈[W1,p(Ω)]2 with p>2 and u0,v0≩0. Then the problem (1.3) has a unique global classical solution (u,v)∈[C0(ˉΩ×[0,∞))∩C2,1(ˉΩ×(0,∞))]2 satisfying u,v>0 for all t>0. Furthermore there exists a constant C>0 independent of t such that
‖u(⋅,t)‖L∞+‖v(⋅,t)‖W1,∞≤C. |
For the global stability, except for the hypotheses (H1)-(H2), we also need the following hypothesis on the compound function
ϕ(v):=v(a2−b2v)F(v). | (1.4) |
(H3) The function ϕ(v) is continuously differentiable on (0,∞), ϕ(0)=limv→0ϕ(v)>0 and ϕ′(v)≤0 for any v≥0.
Remark 1.1. We remark that the hypothesis (H3) is not stringent, and can be satisfied by many forms by imposing some conditions on the parameters if needed. For example, if F(v) is of Holling type Ⅰ or Holling type Ⅱ with a2≤b2λ, then (H3) is automatically satisfied. In general, if (H3) is violated, pattern formations such as periodic orbits or non-constant steady state may arise (see [40]).
Another relevant question is whether the interacting predator-prey population will arrive at the coexistence, exclusion or extinction eventually, which is always an important topic in population dynamics. One can easily compute that the system (1.3) has four possible steady states:
(us,vs)=(0,0)or(0,a2b2)or(a1b1,0)or(u∗,v∗), |
where (u∗,v∗) satisfies
u∗=v∗(a2−b2v∗)F(v∗),αF(v∗)=b1u∗−a1. | (1.5) |
By constructing suitable Lyapunov functionals, we can obtain the following global stability of the coexistence steady state (u∗,v∗) and semi-trivial steady state (a1b1,0). In the context, we define
K:=max{a2b2,‖v0‖L∞}. | (1.6) |
Theorem 1.2 (Global stability). Let Ω⊂R2 be a bounded domain with smooth boundary and the hypotheses (H1)-(H2) hold. Assume (u0,v0)∈[W1,p(Ω)]2 with p>2 and u0,v0≩0 and (u,v) is the solution of (1.3) obtained in Theorem 1.1.
(1) If (H3) holds and the parameters satisfy
max0≤v≤Ku∗F2(v)|d′(v)|24αF(v∗)F′(v)d(v)≤1, |
then
‖u−u∗‖L∞+‖v−v∗‖L∞→0ast→∞, |
and (u∗,v∗) satisfies (1.5), where K is defined in (1.6). Moreover, there exist some positive constants σ,T0 and C independent of t such that
‖u(⋅,t)−u∗‖L∞+‖v(⋅,t)−v∗‖L∞≤Ce−σt,t>T0. |
(2) If the parameters satisfy
limv→0F(v)vexists, minˉΩF(v)v≥a2b1a1,anda1a2|d′(v)|22b1b2d(v)≤1ξ1, | (1.7) |
then
‖u−a1b1‖L∞+‖v‖L∞→0ast→∞, |
where ξ1 and K0 are defined by ξ1=1α+b1b2K20α2 and K0=maxˉΩF(v)v respectively. Moreover, there exist some positive constants T1 and C independent of t such that
‖u(⋅,t)−a1b1‖L∞+‖v(⋅,t)‖L∞≤C1+t,t>T1. |
Remark 1.2. For the semi-trivial steady state (0,a2b2), in the special case of F(v)=v (Holling type Ⅰ), we have showed that it is linear unstable in section 5. For more general F, the globally stability for (0,a2b2) is nontrivial and has to be left open in the current paper.
In the proof of global existence, the method used in [31] was based on a priori estimates for the energy functional ∫Ωulnudx+∫Ω|∇v|2dx to attain the L2 estimates of the solutions. However, such a method of a priori estimates is only applicable for the case where the motility function d(v) is constant. Therefore, the method in [31] is not adaptable to the model (1.3). In this paper, we first derive the L2 estimates for |∇v| and then directly establish the L2 estimates for u by the Gagliardo-Nirenberg inequality and regularity lemmas, in which we also need to prove the boundedness of ∫t+τt∫Ωu2dxds and ∫t+τt∫Ω|Δv|2dxds. In the proof, we do not use the property of the self-adjoint realisation of −Δ+δ (see [29]). Finally, we derive the boundedness for u by the Moser iteration technique.
If d(v) is constant, the system (1.3) has been studied from many aspects as we mentioned above. If d(v) is non-constant as considered in this paper, we find that the system (1.3) can generate pattern formation as presented in section 5 under the condition d′(v)>0. The pattern formation is obviously different from the ones in [29], in which the authors studied the case d′(v)<0. In our case a1>0, the corresponding predator is the generalist predator, the pattern formation may not occur when d′(v)≤0 by linear instability analysis and numerical simulations.
The paper is organized as follows. In section 2, we present the local existence theorem with some preliminary results, and we derive the boundedness of ‖v(⋅,t)‖L∞ and ‖∇v(⋅,t)‖L2. In section 3, we derive the boundedness of ‖u(⋅,t)‖L∞ by the technique of Moser iteration. In section 4, we construct suitable Lyapunov functionals and use the LaSalle invariance principle to prove the global stability and convergence rate stated in Theorem 1.2. In section 5, we further explore time-periodic patterns by linear instability analysis and numerical simulations.
In the sequel, we shall use C or Ci to denote a positive generic constant which may vary in the context. Without confusion, the integration variables x and t will be omitted, for instance ∫a0∫Ωf(x,t)dxdt will be abbreviated as ∫a0∫Ωf(x,t). Often ‖f‖Lp(Ω) will be written as ‖f‖Lp. The existence and uniqueness of local solutions to (1.3), which can be readily proved by the Amann theorem [32,33] or the well-established fixed pointed argument together with the parabolic regularity theory [12].
Lemma 2.1 (Local existence). Let the assumptions in Theorem 1.1 hold. Then there exists a constant Tmax∈(0,∞] such that the problem (1.3) admits a unique classical solution
(u,v)∈[C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax))]2 |
satisfying u,v>0 for all t>0. Moreover,
ifTmax<∞, then lim supt↗Tmax(‖u(⋅,t)‖L∞+‖v(⋅,t)‖W1,∞)=∞. | (2.1) |
Proof. Denote z=(u,v). Then the system (1.3) can be written as
{zt=∇⋅(P(z)∇z)+Q(z),x∈Ω,t>0,∂z∂ν=0,x∈∂Ω,t>0,z(⋅,0)=(u0,v0),x∈Ω, |
where
P(z)=(d(v)d′(v)u01),Q(z)=(u(a1−b1u+αF(v))v(a2−b2v)−uF(v)). |
Since the given initial value of (u0,v0) are nonnegative and satisfy 0≤(u0,v0)∈[W1,p(Ω)]2 with p>2, and hence the matrix P(z) is positive definite at t=0. This means that the system (1.3) is uniformly parabolic. Then the application of [33,Theorem 7.3] yields a Tmax>0 such that the system (1.3) possesses a unique solution (u,v)∈[C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax))]2.
Next, we prove the positivity of u and v. To this end, we rewrite the first equation of (1.3) as
ut=d(v)Δu+(d′(v)∇v+d′(v)∇v)⋅∇u+d″(v)u|∇v|2+d′(v)uΔv+u(a1−b1u+αF(v)). | (2.2) |
Applying the strong maximum principle to (2.2) with the Neumann boundary condition deduces that u>0 for all (x,t)∈Ω×(0,Tmax) due to the fact u0≩0. In a similar way, we can prove v>0 for any (x,t)∈Ω×(0,Tmax) by using the second equation in (1.3). In addition, since P(z) is an upper triangular matrix, the blow-up criterion (2.1) follows from [35,Theorem 5.2] directly. Then the proof of Lemma 2.1 is completed.
Lemma 2.2. Let the assumptions in Theorem 1.1 hold. Then the solution of (1.3) satisfies
‖v(⋅,t)‖L∞≤K | (2.3) |
for all t>0, where K is defined in (1.6), and
limsupt→∞v(⋅,t)≤a2b2 for all x∈ˉΩ. | (2.4) |
Proof. The proof is the similar to [31,Lemma 2.2], but for readers' convenience, we list the proof here. Since u,v and F(v) are nonnegative, we can derive from the second equation of (1.3) that
{vt−Δv=−uF(v)+v(a2−b2v)≤v(a2−b2v),x∈Ω,t>0,∂v∂ν=0,x∈∂Ω,t>0,v(x,0)=v0(x),x∈Ω. | (2.5) |
Let ˉv(t) be the solution of the following ODE problem
{dˉv(t)dt=ˉv(t)(a2−b2ˉv(t)),t>0,ˉv(0)=‖v0‖L∞. | (2.6) |
Then we obtain from (2.6) that ˉv(t)≤K:=max{a2b2,‖v0‖L∞}. It is obvious that ˉv(t) is one of the super-solution of the following PDE problem
{Vt−ΔV=V(a2−b2V),x∈Ω,t>0,∂V∂ν=0,x∈∂Ω,t>0,V(x,0)=v0(x),x∈Ω, | (2.7) |
and therefore, we have
0<V(x,t)≤ˉv(t)for all (x,t)∈ˉΩ×(0,∞), | (2.8) |
where we have used the strong maximum principle to derive V>0. Combining (2.5), (2.7) with (2.8) and employing the comparison principle, one has
0<v(x,t)≤V(x,t)≤ˉv(t)≤Kfor all (x,t)∈ˉΩ×(0,∞). | (2.9) |
This proves (2.3).
In addition, since v(a2−b2v)<0 for v>K, we can further deduce from (2.6) that
limsupt→∞ˉv(t)≤K for all x∈ˉΩ, |
which together with (2.9) gives (2.4).
Lemma 2.3. Let the assumptions in Theorem 1.1 hold. Then the solution of (1.3) satisfies
∫Ωudx≤C,for allt∈(0,Tmax). | (2.10) |
Moreover, one has
∫t+τt∫Ωu2dxds≤C,for allt∈(0,˜Tmax), | (2.11) |
where
τ:=min{1,Tmax2}and˜Tmax:={Tmax−τ,ifTmax<∞,∞,ifTmax=∞. | (2.12) |
Proof. Multiplying the second equation of (1.3) by α and adding the result into the first equation of (1.3), then integrating the result over Ω, we obtain
ddt∫Ω(u+αv)=∫Ωu(a1−b1u)+∫Ωαv(a2−b2v), |
which implies that
ddt∫Ω(u+αv)dx+∫Ω(u+αv)dx+b12∫Ωu2dx=∫Ωu(a1+1−b12u)dx+∫Ωαv(a2+1−b2v)dx≤((a1+1)22b1+α(a2+1)24b2)|Ω|. | (2.13) |
Applying the Grönwall inequality to (2.13) yields
∫Ω(u+αv)dx≤C, | (2.14) |
it follows that (2.10) is valid. Then integrating (2.13) over (t,t+τ), and using (2.14) to obtain
b12∫t+τt∫Ωu2dxds≤∫Ω(u+αv)dx+((a1+1)22b1+α(a2+1)24b2)|Ω|≤C, |
which implies (2.11).
By Lemma 2.3, we establish the estimates for ‖∇v‖L2 and ∫t+τt∫Ω|Δv|2dxds.
Lemma 2.4. Let the assumptions in Theorem 1.1 hold. Then there exists a constant C>0 independent of t such that
‖∇v‖L2≤C,for allt∈(0,Tmax), | (2.15) |
and
∫t+τt∫Ω|Δv|2dxds≤C,for allt∈(0,˜Tmax), | (2.16) |
where τ and ˜Tmax are defined in (2.12).
Proof. Multiplying the second equation of (1.3) by −Δv, integrating the result in Ω, using the assumption (H2) and the boundedness of v (see (2.3)), we have
12ddt∫Ω|∇v|2dx+∫Ω|Δv|2dx=∫ΩuF(v)Δvdx−∫Ωv(a2−b2v)Δvdx,≤12∫Ω|Δv|2dx+∫Ωu2F2(v)dx+∫Ωv2(a2−b2v)2dx≤12∫Ω|Δv|2dx+F2(K)∫Ωu2dx+C1, |
which implies
ddt∫Ω|∇v|2dx+∫Ω|Δv|2dx≤2F2(K)∫Ωu2dx+2C1, | (2.17) |
where K is defined in (1.6).
Applying the Gagliardo-Nirenberg inequality, [29,Lemma 2.5] and noting the fact ‖v‖L2≤K|Ω|12 yields
∫Ω|∇v|2dx=‖∇v‖2L2≤C2(‖Δv‖L2‖v‖L2+‖v‖2L2)≤12‖Δv‖2L2+C3. | (2.18) |
Substituting (2.18) into (2.17), one has
ddt∫Ω|∇v|2dx+∫Ω|∇v|2dx+12∫Ω|Δv|2dx≤2F2(K)∫Ωu2dx+C4, | (2.19) |
which together with (2.11) yields (2.15).
Then integrating (2.19) over (t,t+τ), we obtain (2.16).
In this section, we prove the boundedness of ‖u‖L∞ by the technique of Moser iteration. To the end, we first prove the boundedness of ‖u‖L2.
Lemma 3.1. Let the assumptions in Theorem 1.1 hold. Then there exists a constant C>0 independent of t such that
‖u(⋅,t)‖L2≤Cfor allt∈(0,Tmax). | (3.1) |
Proof. Multiplying the first equation in (1.3) by 2u and integrating the result with respect to x in Ω, one has
ddt∫Ωu2+2∫Ωd(v)|∇u|2dx+2b1∫Ωu3=−2∫Ωud′(v)∇u⋅∇v+2a1∫Ωu2+2α∫Ωu2F(v)dx. | (3.2) |
By the assumptions in (H1), (H2), and (2.3), we know that d(v)∈C3 and 0<v≤K, where K is defined in (1.6). Therefore, one has d(v)≥C1, 0<F(v)≤F(K) and |d′(v)|≤C2, then it follows from (3.2) that
ddt∫Ωu2+2C1∫Ω|∇u|2dx+2b1∫Ωu3≤2C2∫Ωu|∇u||∇v|dx+a1∫Ωu2dx+2αF(K)∫Ωu2dx≤C1∫Ω|∇u|2dx+C22C1∫Ωu2|∇v|2dx+2(a1+αF(K))∫Ωu2dx≤C1‖∇u‖2L2+C22C1‖u‖2L4‖∇v‖2L4+2(a1+αF(K))‖u‖2L2. | (3.3) |
Applying the Galiardo-Nirenberg inequality, one has
‖u‖2L4≤C3(‖∇u‖L2‖u‖L2+‖u‖2L2) | (3.4) |
and
‖∇v‖2L4≤C4(‖Δv‖L2‖∇v‖L2+‖∇v‖2L2)≤C5(‖Δv‖L2+1), | (3.5) |
where we have used the boundedness of ‖∇v‖L2 (see (2.15)) and [29,Lemma 2.5]. Combining (3.4) with (3.5) and applying the Young inequality yields
C22C1‖u‖2L4‖∇v‖2L4≤C6(‖∇u‖L2‖u‖L2+‖u‖2L2)(‖Δv‖L2+1)≤C6‖∇u‖L2‖u‖L2‖Δv‖L2+C6‖∇u‖L2‖u‖L2+C6‖u‖2L2‖Δv‖L2+C6‖u‖2L2≤C1‖∇u‖2L2+C26C1‖u‖2L2‖Δv‖2L2+C7‖u‖2L2. | (3.6) |
Then substituting (3.6) into (3.3), we obtain
ddt‖u‖2L2≤C26C1‖u‖2L2‖Δv‖2L2+(C7+2αF(K))‖u‖2L2≤C8‖u‖2L2(‖Δv‖2L2+1). | (3.7) |
For any t∈(0,Tmax), by (2.10), there exists a nonnegative t0∈((t−τ)+,t) with τ=min{1,12Tmax} such that
∫Ωu2(x,t0)dx≤C9 | (3.8) |
and by (2.16), one has
∫t0+τt0∫Ω|Δv(x,s)|2dxds≤C10, for all t0∈(0,˜Tmax). | (3.9) |
Then integrating (3.7) on (t0,t), and applying (3.8), (3.9) and the fact t≤t0+τ≤t0+1, we have
‖u(⋅,t)‖2L2≤‖u(⋅,t0)‖2L2eC8∫tt0(‖Δv(⋅,s)‖2L2+1)ds≤C11, |
which indicates (3.1) and thus completes the proof.
To derive the boundedness for ‖u‖L∞, we need the following regularity lemma.
Lemma 3.2. ([37]) Assume that Ω⊂Rn is a bounded domain with smooth boundary. Suppose that y(x,t)∈C2,1(ˉΩ×(0,Tmax)) is the solution of
{yt=Δy−y+ϕ(x,t),x∈Ω,t∈(0,Tmax),∂y∂ν=0,x∈∂Ω,t∈(0,Tmax),y(x,0)=y0(x)∈C0(ˉΩ), |
where ϕ(x,t)∈L∞((0,Tmax);Lp(Ω)). Then there exists a constant C>0 such that
‖y(⋅,t)‖W1,q≤C for all t∈(0,Tmax) |
with
q∈{[1,npn−p),if p≤n,[1,∞],if p>n. |
We will prove the boundedness of u through the Moser iteration procedure.
Lemma 3.3. Let the assumptions in Theorem 1.1 hold. Then there exists a positive constant C independent of t such that
‖u(⋅,t)‖L∞+‖∇v(⋅,t)‖L∞≤Cfor allt∈(0,Tmax). | (3.10) |
Proof. Multiplying the first equation of the system (1.3) by up−1 with p≥2, and integrating the resulting equation by parts, we derive
1pddt∫Ωupdx+(p−1)∫Ωd(v)up−2|∇u|2dx+b1∫Ωup+1dx−a1∫Ωupdx=−(p−1)∫Ωd′(v)up−1∇u⋅∇vdx+α∫ΩupF(v)dx. | (3.11) |
Noting the fact d(v)≥C1, 0<F(v)≤F(K) and |d′(v)|≤C2, and using the Young inequality, we obtain from (3.11) that
1pddt∫Ωupdx+(p−1)C1∫Ωup−2|∇u|2dx+∫Ωupdx≤(p−1)C2∫Ωup−1|∇u||∇v|dx+(a1+1+αF(K))∫Ωupdx≤(p−1)C12∫Ωup−2|∇u|2dx+C22(p−1)2C1∫Ωup|∇v|2dx+(a1+1+αF(K))∫Ωupdx, |
which gives
ddt∫Ωupdx+p∫Ωupdx+2(p−1)C1p∫Ω|∇up2|2dx≤C22p(p−1)2C1∫Ωup|∇v|2dx+p(a1+1+αF(K))∫Ωupdx | (3.12) |
for all t∈(0,Tmax) and p≥2. By Lemma 3.1, we have ‖u(⋅,t)‖L2≤C3, and thus we derive ‖∇v(⋅,t)‖L4≤C4 from Lemma 3.2. Then applying the Gagliardo-Nirenberg inequality and the Hölder inequality yields
C22p(p−1)2C1∫Ωup|∇v|2dx≤C22p(p−1)2C1(∫Ωu2pdx)12(∫Ω|∇v|4dx)12≤C22C24p(p−1)2C1‖up2‖2L4≤C5(‖∇up2‖2(1−1p)L2‖up2‖2pL4p+‖up2‖2L4p)≤C3C5‖∇up2‖2(1−1p)L2+Cp3C5≤(p−1)C1p‖∇up2‖2L2+C1p(C3C5C1)p+Cp3C5, | (3.13) |
and
p(a1+1+αF(K))∫Ωupdx=p(a1+1+αF(K))‖up2‖2L2≤p(a1+1+αF(K))(‖∇up2‖2(1−2p)L2‖up2‖4pL4p+‖up2‖2L4p)≤C23C6‖∇up2‖2(1−2p)L2+Cp3C6≤(p−1)C1p‖∇up2‖2L2+C1p(C23C6d(0))p+Cp3C6. | (3.14) |
Substituting (3.13) and (3.14) into (3.12), we obtain
ddt∫Ωupdx+p∫Ωupdx≤C7, |
by the Grönwall inequality, we derive
‖u(⋅,t)‖pLp≤e−pt‖u0‖pLp+C7p(1−e−pt)≤‖u0‖pLp+C7p. |
Let p=4 in the above inequality and apply Lemma 3.2 again. We obtain that there exists a constant C7>0 such that
‖∇v(⋅,t)‖L∞≤C7. |
Then using the Moser iteration procedure (see [34]), one derives (3.10) and thus proves Lemma 3.3.
Proof of Theorem 1.1. Theorem 1.1 is a consequence of Lemma 2.1, Lemma 2.2 and Lemma 3.3.
In this section, we will investigate the asymptotical behavior of solutions solving system (1.3) and prove Theorem 1.2 based on Lyapunoval functional method along with Barbălat's lemma.
Let us first recall the basic Barbălat's lemma.
Lemma 4.1. (Barbălat's Lemma [36]) Suppose that h:[1,∞)→R is a uniformly continuous function such that limt→∞∫t1h(s)ds exists, then limt→∞h(t)=0.
Now we give the uniform estimates of the global solution.
Lemma 4.2. Let (u,v) be the unique global bounded classical solution of (1.3) given by Theorem 1.1. Then for any given 0<α<1, there exists a constant C(α)>0 such that
‖u‖C2+α,1+α2(ˉΩ×[1,∞))+‖v‖C2+α,1+α2(ˉΩ×[1,∞))≤C(α). | (4.1) |
Proof. This proof is based on the standard regularity for parabolic equations. For readers' convenience, we sketch the proof here. Due to the boundedness of (u,v), applying the interior Lp estimate [39] to (1.3), we derive that
‖u‖W2,1p(Ω×[i+14,i+3])+‖v‖W2,1p(Ω×[i+14,i+3])≤C1,∀i≥0. | (4.2) |
Using the Sobolev embedding theorem and (4.2), we derive
‖u‖C1+α,1+α2(ˉΩ×[14,∞))+‖v‖C1+α,1+α2(ˉΩ×[14,∞))≤C2. | (4.3) |
Applying (4.3) and the Schauder estimate [38] to the second equation of (1.3), we obtain
‖v‖C2+α,1+α2(ˉΩ×[i+13,i+3])≤C3,∀i≥0, | (4.4) |
which implies
‖v‖C2+α,1+α2(ˉΩ×[13,+∞))≤C4. | (4.5) |
Rewrite the first equation in (1.3) as
ut−d(v)Δu−2d′(v)∇v⋅∇u=G(x,t),x∈Ω,t>0, | (4.6) |
where
G(x,t)=(d″(v)|∇v|2+d′(v)Δv)u+u(a1−b1u+αF(v)). |
Due to (4.3) and (4.4), we see that
‖G‖Cα,α2(ˉΩ×[i+13,i+3])≤C5,∀i≥0. |
Applying the Schauder estimate to (4.6) gives ‖u‖C2+α,1+α2(ˉΩ×[i+1,i+3])≤C6 for all i≥0. Thus
‖u‖C2+α,1+α2(ˉΩ×[1,+∞))≤C7. | (4.7) |
Then (4.1) follows from (4.5) and (4.7). This completes the proof of Lemma 4.2.
Next we shall prove the global stability of the coexistence steady state (u∗,v∗) by constructing the following Lyapunov functional:
V(u(t),v(t))=V(t)=1α∫Ω(u−u∗−u∗lnuu∗)dx+∫Ω∫vv∗F(s)−F(v∗)F(s)dsdx, | (4.8) |
where (u∗,v∗) satisfies (1.5).
Lemma 4.3. Let Ω⊂R2 be a bounded domain with smooth boundary and the hypotheses (H1)-(H3) hold. Assume (u0,v0)∈[W1,p(Ω)]2 with p>2 and u0,v0≩0. If (u,v) is the solution of (1.3) obtained in Theorem 1.1, and the parameters satisfy
max0≤v≤Ku∗F2(v)|d′(v)|24αF(v∗)F′(v)d(v)≤1, |
then
‖u−u∗‖L∞+‖v−v∗‖L∞→0ast→∞, | (4.9) |
where (u∗,v∗) satisfies (1.5), and K is defined in (1.6).
Proof. First by a similar argument as [31,Lemma 4.3], we deduce that
V(t)≥0 for all u,v≥0. |
We compute the derivative of V(t) by using the system (1.3) to derive
ddtV(t)=1α∫Ω(1−u∗u)utdx+∫ΩF(v)−F(v∗)F(v)vtdx=−u∗α∫Ωd(v)|∇u|2u2dx−u∗α∫Ωd′(v)∇u⋅∇vudx−F(v∗)∫ΩF′(v)|∇vF(v)|2dx⏟I1+1α∫Ω(u−u∗)(a1−b1u+αF(v))dx+∫Ω(F(v)−F(v∗))(v(a2−b2v)F(v)−u)⏟I2. | (4.10) |
Then I1 can be rewritten as
I1=−∫ΩXAXT, |
where XT denotes the transpose of X:=(∇u,∇v), and
A=(u∗d(v)αu2d′(v)u∗2αud′(v)u∗2αuF(v∗)F′(v)|F(v)|2). |
It is easy to check that the matrix A is nonnegative definite if and only if
u∗F2(v)|d′(v)|24αF(v∗)F′(v)d(v)≤1, |
and thus
I1≤0. | (4.11) |
Next, we compute I2. By (1.5), we see that
a1−b1u∗+αF(v∗)=0andv∗(a2−b2v∗)F(v∗)−u∗=0, |
we obtain from (H2) and (H3) that there exists a constant T1>0 such that
0<12F′(v∗)≤F′(ξ1)≤2F′(v∗),ϕ′(v∗)≤ϕ′(ξ2)≤12ϕ′(v∗)<0. |
Therefore, there exists a constant C1>0 such that
I2=−b1α∫Ω(u−u∗)2dx+∫Ω(F(v)−F(v∗))(ϕ(v)−ϕ(v∗))dx≤−b1α∫Ω(u−u∗)2dx+∫ΩF′(ξ1)ϕ′(ξ2)(v−v∗)2dx≤−C∫Ω[(u−u∗)2+(v−v∗)2]dx:=−C1E(t), | (4.12) |
where ϕ(v)=v(a2−b2v)F(v) is defined in (1.4), ξ1 and ξ2 are lying between v and v∗. Combining (4.10), (4.11) with (4.12) gives
ddtV(t)≤−C1E(t) | (4.13) |
with E(t)=∫Ω[(u−u∗)2+(v−v∗)2]dx.
Since V(t)≥0, we have
∫∞1E(t)dt≤1C1V(1)<∞. |
It follows from the regularity of u,v that E(t) is uniformly continuous in [1,∞). An application of Lemma 4.1 yields
E(t)=∫Ω[(u−u∗)2+(v−v∗)2]dx→0 as t→∞. | (4.14) |
By Lemma 4.2, we derive that u(⋅,t) and v(⋅,t) are bounded for t>1 in the space W1,∞(Ω). Applying the Gagliardo-Nirenberg inequality
‖ϕ‖∞≤c‖ϕ‖nn+2W1,∞(Ω)‖ϕ‖2n+22,∀ϕ∈W1,∞(Ω) |
to u−u∗ and v−v∗, respectively, we can obtain (4.9) from (4.14). This completes the proof of Lemma 4.3.
Lemma 4.4. Suppose that the conditions of Lemma 4.3 hold. Then there exist some positive constants σ,T0 and C independent of t such that
‖u(⋅,t)−u∗‖L∞+‖v(⋅,t)−v∗‖L∞≤Ce−σt,t>T0. | (4.15) |
Proof. By Lemma 4.3, we have
‖u−u∗‖L∞→0ast→∞, |
then applying L'Hôpital's rule to derive
limu→u∗u−u∗−u∗lnuu∗(u−u∗)2=12u∗. |
Therefore, we obtain from the continuity that there exists a constant t1>0 such that
14u∗∫Ω(u−u∗)2≤∫Ω(u−u∗−u∗lnuu∗)≤1u∗∫Ω(u−u∗)2,t>t1. | (4.16) |
In addition, by Lemma 4.1 in [31], we find that there exist some constants C1,C2>0 and t2>0 such that
C1∫Ω(v−v∗)2≤∫Ω∫vv∗F(s)−F(v∗)F(s)ds≤C2∫Ω(v−v∗)2,t>t2. | (4.17) |
Combining (4.16), (4.17), and the definition of V(t) in (4.8), we conclude that there exist some positive constants C3,C4 and t3 such that for all t>t3
C3E(t)≤V(t)≤C4E(t). |
This, together with (4.13) yields a constant C5>0 such that
ddtV(t)≤−C5V(t),t>t3, |
therefore, we derive that there exist some constants C6>0 and C7>0 such that
‖u−u∗‖2L2+‖v−v∗‖2L2≤C6e−C7t. | (4.18) |
To finish the proof, we need the L∞ estimates for u−u∗ and v−v∗. By Lemma 4.2 and the Gagliardo-Nirenberg inequality, we have
‖u−u∗‖L∞≤C8(‖∇u‖23L4‖u−u∗‖13L2+‖u−u∗‖L2)≤C9‖u−u∗‖13L2, | (4.19) |
by a similar argument, one has
‖v−v∗‖L∞≤C10‖u−u∗‖13L2, |
which, combined with (4.18) and (4.19), gives (4.15). Therefore, we complete the proof of Lemma 4.4.
Lemma 4.5. Let Ω⊂R2 be a bounded domain with smooth boundary and the hypotheses (H1)-(H2) hold. Assume (u0,v0)∈[W1,p(Ω)]2 with p>2 and u0,v0≩0. If (u,v) is the solution of (1.3) obtained in Theorem 1.1, and the parameters satisfies
limv→0F(v)v exists, minˉΩF(v)v≥a2b1a1, and a1a2(d′(v))22b1b2d(v)≤1ξ1, | (4.20) |
then
‖u−a1b1‖L∞+‖v‖L∞→0ast→∞, | (4.21) |
where K is defined in (1.6), ξ1 and K0 are defined by ξ1=1α+b1b2K20α2 and K0=maxˉΩF(v)v respectively.
Next we prove the global stability of the generalist predator only steady state (a1b1,0) by constructing the following Lyapunov functional:
V1(u(t),v(t))=V1(t)=ξ1∫Ω(u−a1b1−a1b1lnua1b1)dx+∫Ωvdx+b24a2∫Ωv2dx. |
Proof. It is obvious that V1(t)≥0 from the definition. Next we compute the derivative of V1(t) by using the system (1.3) and (4.20) to obtain
ddtV1(t)=ξ1∫Ω(1−a1b1u)utdx+∫Ωvtdx+b22a2∫Ωvvtdx=−ξ1a1b1∫Ωd(v)|∇u|2u2−ξ1a1b1∫Ωd′(v)∇u⋅∇vu−ξ1b1∫Ω(u−a1b1)2dx+αξ1∫Ω(u−a1b1)F(v)dx+a2∫Ωvdx−b22∫Ωv2dx−∫Ω(u−a1b1)F(v)dx−a1b1∫ΩF(v)dx−b22a2∫Ω|∇v|2dx−b222a2∫Ωv3dx−b22a2∫ΩuvF(v)dx≤−ξ1a1b1∫Ωd(v)|∇u|2u2−ξ1a1b1∫Ωd′(v)∇u⋅∇vu−b22a2∫Ω|∇v|2dx⏟J1−ξ1b1∫Ω(u−a1b1)2dx−b22∫Ωv2dx+K0|αξ1−1|∫Ω(u−a1b1)vdx⏟J2. | (4.22) |
Then J1 and J2 can be rewritten as
J1=−∫ΩX1A1XT1dx,J2=−∫ΩY1B1YT1dx, |
where XT1 and YT1 denotes the transpose of X1:=(∇uu,∇v) and Y1:=(u−a1b1,v), and
A1=(a1ξ1d(v)b1a1ξ1d′(v)2b1a1ξ1d′(v)2b1b22a2), |
and
B1=(b1ξ1−K0|αξ1−1|2−K0|αξ1−1|2b22), |
By the definition of ξ1 and K0 defined in Lemma 4.5, and (4.20), we derive that
J1≤0 |
and
J2≤−C(∫Ω(u−a1b1)2dx+∫Ωv2dx):=−CE1(t). |
It follows that
ddtV1(t)≤−CE1(t) | (4.23) |
with E1(t)=∫Ω(u−a1b1)2dx+∫Ωv2dx.
By a similar argument as Lemma 4.3, we obtain (4.21) and thus completes the proof of Lemma 4.5.
By a similar argument as Lemma 4.4, we can derive the following decay rate estimates.
Lemma 4.6. Suppose that the conditions in Lemma 4.5 hold. Then there exist some positive constants T1 and C independent of t such that
‖u(⋅,t)−a1b1‖L∞+‖v(⋅,t)‖L∞≤C1+t,t>T1. | (4.24) |
Proof. Since
V1(t)≤CE121(t), |
then combining this with (4.23), we obtain
ddtV1(t)≤−CV21(t). |
Solving this ordinary differential inequality, for sufficiently large t, we arrive at
V1(t)≤C1+t. |
Using the same argument as in the proof of Lemma 4.4, we readily get (4.24) and thus complete the proof of Lemma 4.6.
In this section, we shall study the possible pattern formation generated by the system (1.3). As a typical example, we consider the case Holling type Ⅰ of F(v)=v.
To begin with, we first consider the linear stability of the system (1.3). We linearise the system (1.3) at an equilibrium (us,vs) and write the linearised system as
{Φt=AΔΦ+BΦ,x∈Ω,t>0,(ν⋅∇)Φ=0,x∈∂Ω,t>0,Φ(x,0)=(u0−us,v0−vs)T,x∈Ω, | (5.1) |
where T denotes the transpose and
Φ=(u−usv−vs),A=(d(vs)usd′(vs)01), |
as well as
B=(a1−2b1us+αvsαus−vs−us+a2−2b2vs). |
Let Wk(x) be the eigenfunction of the following eigenvalue problem:
{ΔWk(x)+k2Wk(x)=0,∂Wk∂ν(x)=0, |
the allowable wavenumbers k are discrete in a bounded domain, for instance, if Ω=(0,l), then k=nπl,n=0,1,2,⋯. Since the system (5.1) is linear, the solution of Φ(x,t) has the form of
Φ(x,t)=∑k≥0ckeρtWk(x), | (5.2) |
where ρ is the temporal eigenvalue and ck(k≥0) are determined by the Fourier expansion of the initial conditions in terms of Wk(x). Substituting (5.2) into (5.1) yields
ρWk(x)=−k2AWk(x)+BWk(x), |
which indicates that ρ is the eigenvalue of the matrix Mk with
Mk=(−k2d(vs)+a1−2b1us+αvs−k2usd′(vs)+αus−vs−k2−us+a2−2b2vs). |
Then the eigenvalue ρ(k2) satisfies
ρ2+a(k2)ρ+b(k2)=0, |
where
a(k2)=(d(vs)+1)k2−a1+2b1us−αvs+us−a2+2b2vs |
and
b(k2)=(−k2d(vs)+a1−2b1us+αvs)(−k2−us+a2−2b2vs)−vs(k2usd′(vs)−αus). |
Next we proceed to consider the stability of equilibria (0,0), (a1b1,0), (0,a2b2) and (a2α+a1b2α+b1b2,a2b1−a1α+b1b2) in the presence of spatial structure.
● For the steady state (0,0), it can be easily check that ρ satisfies
(ρ−a1)(ρ+k2−a2)=0. |
At least one of the roots is positive, therefore, (0,0) is always linearly unstable.
● For the steady state (a1b1,0), it can be easily check that ρ satisfies
(ρ+k2d(0)+a1)(ρ+k2+us−a2)=0. |
Therefore if a2=us, (a1b1,0) is marginally stable. If a2>us, then there exists a k such that one of the root is negative and the other is positive, and (a1b1,0) is linearly unstable.
● For the steady state (0,a2b2), ρ satisfies
(ρ+k2d(vs)−a1−αvs)(ρ+k2+a2)=0. |
Therefore there exists a k such that one of the root is negative and the other is positive, and (0,a2b2) is linearly unstable.
● For the coexistence steady state (us,vs)=(a2α+a1b2α+b1b2,a2b1−a1α+b1b2), ρ satisfies
ρ2−(M1+M4)ρ+M1M4−M2M3=0, |
with M1=−k2d(vs)−b1us,M2=−k2usd′(vs)+αus,M3=−vs and M4=−k2−b2vs, where we have used the identity
a1−b1us+αvs=0anda2−b2vs−us=0. |
Since M1+M4<0, (us,vs) is linearly unstable iff
M1M4−M2M3<0. |
One root is negative and the other is positive. That is, we need the condition
(k2d(vs)+b1us)(k2+b2vs)+vs(αus−k2usd′(vs))<0. |
which indicates that
d(vs)k4+(b2vsd(vs)+b1us−usvsd′(vs))k2+(b1b2+α)usvs<0. | (5.3) |
We conclude that a steady-state bifurcation may occur if
usvsd′(vs)−b2vsd(vs)−b1us>2√(b1b2+α)d(vs)usvs |
and there are allowable wavenumbers k such that
k−1<k2<k+1, |
where k±1=usvsd′(vs)−b2vsd(vs)−b1us)±√(b2vsd(vs)+b1us−usvsd′(vs))2−4(b1b2+α)d(vs)usvs2d(vs).
We remark here that when d′(v)≤0, the inequality (5.3) can not be hold for any k. Therefore, the coexistence steady state (us,vs) is linearly stable if d′(v)≤0.
In this subsection, we take some examples to present the periodic patterns.
According to the condition (5.3) and the linear stability analysis in subsection 5.1, we fix the value of the parameters in all simulations as follows:
a1=b1=1,a2=b2=α=2,d(v)=e20vandl=10. | (5.4) |
Then we obtain the possible steady states are (us,vs)=(32,14) or (1,0) or (0,1).
The numerical simulations of patterns are then shown in the following Figures 1–3.
Case 1. (us,vs)=(32,14). In this case, k=nπ10,n=0,1,2,⋯, the condition (5.3) turns into
e5k4+(32−7e5)k2+32<0, |
it can be checked that the above inequality is valid iff 1≤n≤8.
The numerical spatial-temporal patterns generated by the model (1.3) under the condition (5.4) are plotted in Figure 1(a), where we observe the spatially inhomogeneous temporal periodic patterns arising from the vicinity of equilibrium (32,14). The time distributions of predators and prey at a fixed space position plotted in Figure 1(b) show that predators and prey are periodic. We also plot the spatial distributions of predators and prey at a fixed time in Figure 1(b).
Case 2. (us,vs)=(1,0). In this case, from the above linear analysis, we know that ρ satisfies
(ρ+k2+1)(ρ+k2−1)=0, |
therefore, then one root is positive and the other is negative for the above equation iff n=0,1,2,3.
Similarly, we observe the spatially inhomogeneous temporal periodic patterns arising from the vicinity of equilibrium (1,0). The time distributions of predators and prey at a fixed space position plotted in Figure 2(b) show that predators and prey are periodic. We also plot the spatial distributions of predators and prey at a fixed time in Figure 2(b).
Case 3. (us,vs)=(0,1). In this case, we know that ρ satisfies
(ρ+e20k2−3)(ρ+k2+2)=0, |
therefore, then one root is positive and the other is negative for the above equation iff n=0.
We also can observe the spatially inhomogeneous temporal periodic patterns arising from the vicinity of equilibrium (0,1). The time distributions of predators and prey at a fixed space position plotted in Figure 3(b) show that predators and prey are periodic. We also plot the spatial distributions of predators and prey at a fixed time in Figure 3(b).
In summary, we conclude that the time-periodic patterns have been obtained when d(v) is monotone increasing and satisfies some conditions. While from the linear stability analysis, we know that the coexistence steady state and the semi-trivial steady states are both linearly stable. These results indicate that the motility function d(v) can trigger pattern formation and is a factor inducing the spatial heterogeneity of populations. In addition, if a1<0, (u represents the specialist predator) Jin and Wang in [29] have proved that the semi-trivial steady state (0,a2b2) is linearly stable if αa2b2≤−a1. While our results indicate that when u is a generalist predator (a1>0), the semi-trivial steady state (0,a2b2) is linearly unstable, because the generalist predator u can gain food from other preys.
The authors are grateful to the referee for his/her valuable comments, which greatly improved the exposition of our paper. The research of T. Feng was partially supported by the Special Basic Cooperative Research Programs of Yunnan Provincial Undergraduate Universities Association (Grant No. 2019F001-078, 202101BA070001-132). The research of L. Wu was supported by the 2020 Hong Kong Scholars Program (Project ID P0031250 and Primary Work Programme YZ3S).
The authors declare there is no conflict of interest.
[1] | H. I. Freedman, Deterministic mathematical models in population ecology, volume 57. Marcel Dekker Incorporated, 1980. |
[2] | A. J. Lotka, Elements of mathematical biology, Dover Publications, 1956. |
[3] | A. D Bazykin, Nonlinear dynamics of interacting populations, World Scientific, 1998. https://doi.org/10.1142/2284 |
[4] |
H. I. Freedman, R. M. Mathsen, Persistence in predator-prey systems with ratio-dependent predator influence, Bull. Math. Biol., 55 (1993), 817–827. https://doi.org/10.1016/S0092-8240(05)80190-9 doi: 10.1016/S0092-8240(05)80190-9
![]() |
[5] |
G. F. Gause, N. P. Smaragdova, A. A. Witt, Further studies of interaction between predators and prey, J. Anim. Ecol., 5 (1936), 1–18. https://doi.org/10.2307/1087 doi: 10.2307/1087
![]() |
[6] |
M. P. Hassell, R. M. May, Generalist and specialist natural enemies in insect predator-prey interactions, J. Anim. Ecol., 55 (1986), 923–940. https://doi.org/10.2307/4425 doi: 10.2307/4425
![]() |
[7] |
H.-Y. Jin, Z.-A. Wang, L. Wu, Global dynamics of a three-species spatial food chain model, J. Differ. Equ., 333 (2022), 144–183. https://doi.org/10.1016/j.jde.2022.06.007 doi: 10.1016/j.jde.2022.06.007
![]() |
[8] | J. Maynard-Smith, Models in ecology, Cambridge university press, 1974. |
[9] |
P. Kareiva, G. Odell, Swarms of predators exhibit "preytaxis" if individual predators use area-restricted search, Am. Nat., 130 (1987), 233–270. https://doi.org/10.1086/284707 doi: 10.1086/284707
![]() |
[10] |
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
![]() |
[11] |
C. Liu, X. Fu, L. Liu, X. Ren, C. 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
![]() |
[12] |
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
![]() |
[13] |
M. Ma, R. Peng, Z.-A. 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 doi: 10.1016/j.physd.2019.132259
![]() |
[14] |
J. Smith-Roberge, D. Iron, T. Kolokolnikov, Pattern formation in bacterial colonies with density-dependent diffusion, Eur. J. Appl. Math., 30 (2019), 196–218. https://doi.org/10.1017/S0956792518000013 doi: 10.1017/S0956792518000013
![]() |
[15] |
Z.-A. Wang, L. Y. Wu, Global solvability of a class of reaction-diffusion systems with cross-diffusion, Appl. Math. Lett., 124 (2022), 107699. https://doi.org/10.1016/j.aml.2021.107699 doi: 10.1016/j.aml.2021.107699
![]() |
[16] |
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
![]() |
[17] |
J. Ahn, C. Yoon, Global well-posedness and stability of constant equilibria in parabolic–elliptic chemotaxis systems without gradient sensing, Nonlinearity, 32 (2019), 1327–1351. https://doi.org/10.1088/1361-6544/aaf513 doi: 10.1088/1361-6544/aaf513
![]() |
[18] |
Y. S. Tao, M. Winkler, Effects of signal-dependent motilities in a Keller-Segel-type reaction-diffusion system, Math. Models Methods Appl. Sci., 27 (2017), 1645–1683. https://doi.org/10.1142/S0218202517500282 doi: 10.1142/S0218202517500282
![]() |
[19] |
K. Fujie, J. Jiang, Global existence for a kinetic model of pattern formation with density-suppressed motilities, J. Differ. Equ., 269 (2020), 5338–5378. https://doi.org/10.1016/j.jde.2020.04.001 doi: 10.1016/j.jde.2020.04.001
![]() |
[20] |
K. Fujie, J. Jiang, Comparison methods for a Keller-Segel-type model of pattern formations with density-suppressed motilities, Calc. Var. Partial Differ. Equ., 60 (2021), 37. https://doi.org/10.1007/s00526-021-01943-5 doi: 10.1007/s00526-021-01943-5
![]() |
[21] |
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 doi: 10.1090/proc/15124
![]() |
[22] |
Z. R. Liu, J. Xu, Large time behavior of solutions for density-suppressed motility system in higher dimensions, J. Math. Anal. Appl., 475 (2019), 1596–1613. https://doi.org/10.1016/j.jmaa.2019.03.033 doi: 10.1016/j.jmaa.2019.03.033
![]() |
[23] |
W. B. Lyu, Z.-A. Wang, Global classical solutions for a class of reaction-diffusion system with density-suppressed motility, Electron. Res. Arch., 30 (2022), 995–1035. https://doi.org/10.3934/era.2022052 doi: 10.3934/era.2022052
![]() |
[24] |
H.-Y. Jin, Z.-A. Wang, The Keller-Segel system with logistic growth and signal-dependent motility, Discrete Contin. Dyn. Syst. Ser. B, 26 (2021), 3023–3041. https://doi.org/10.3934/dcdsb.2020218 doi: 10.3934/dcdsb.2020218
![]() |
[25] |
J. P. Wang, M. X. 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
![]() |
[26] |
Z.-A. Wang, On the parabolic-elliptic Keller-Segel system with signal-dependent motilities: a paradigm for global boundedness and steady states, Math. Methods Appl. Sci., 44 (2021), 10881–10998. https://doi.org/10.1002/mma.7455 doi: 10.1002/mma.7455
![]() |
[27] | Z.-A. Wang, J. Xu, On the Lotka-Volterra competition system with dynamical resources and density-dependent diffusion, J. Math. Biol., 82 (2021), Paper No. 7. https://doi.org/10.1007/s00285-021-01562-w |
[28] |
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
![]() |
[29] |
H.-Y. Jin, Z.-A. Wang, Global dynamics and spatio-temporal patterns of predator-prey systems with density-dependent motion, Eur. J. Appl. Math., 32 (2021), 652–682. https://doi.org/10.1017/S0956792520000248 doi: 10.1017/S0956792520000248
![]() |
[30] | K. Nakashima, Y. Yamada, Positive steady states for prey-predator models with cross-diffusion, Adv. Differ. Equ., 1 (1996), 1099–1122. |
[31] |
H.-Y. Jin, Z.-A. Wang, Global stability of prey-taxis systems, J. Differ. Equ., 262 (2017), 1257–1290. https://doi.org/10.1016/j.jde.2016.10.010 doi: 10.1016/j.jde.2016.10.010
![]() |
[32] | H. Amann, Dynamic theory of quasilinear parabolic equations. Ⅱ. reaction-diffusion systems, Diff. Integral Eqns., 3 (1990), 13–75. |
[33] | H. Amann, Nonhomogeneous linear and quasilinear elliptic and parabolic boundary value problems, In Function spaces, differential operators and nonlinear analysis (Friedrichroda, 1992), volume 133 of Teubner-Texte Math., pages 9–126. Teubner, Stuttgart, 1993. |
[34] |
N. D. Alikakos, Lp bounds of solutions of reaction-diffusion equations, Comm. Partial Differ. Equ., 4 (1979), 827–868. https://doi.org/10.1080/03605307908820113 doi: 10.1080/03605307908820113
![]() |
[35] |
H. Amann, Dynamic theory of quasilinear parabolic equations Ⅲ. Global existence, Math. Z., 202 (1989), 219–250. https://doi.org/10.1007/BF01215256 doi: 10.1007/BF01215256
![]() |
[36] | I. Barbălat, Systèmes d'équations différentielles d'oscillations non linéaires, Rev. Math. Pures Appl., 4 (1959), 267–270. |
[37] |
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
![]() |
[38] | O. A. Ladyźenskaja, V. A. Solonnikov, N. N. Ural'ceva, Linear and quasilinear equations of parabolic type, Translated from the Russian by S. Smith. Translations of Mathematical Monographs, Vol. 23. American Mathematical Society, Providence, R.I., 1968. |
[39] | G. M. Lieberman, Second order parabolic differential equations, World Scientific Publishing Co., Inc., River Edge, NJ, 1996. https://doi.org/10.1142/3302 |
[40] |
F. Q. Yi, J. J. Wei, J. P. Shi, Bifurcation and spatiotemporal patterns in a homogeneous diffusive predator-prey system, J. Differ. Equ., 246 (2009), 1944–1977. https://doi.org/10.1016/j.jde.2008.10.024 doi: 10.1016/j.jde.2008.10.024
![]() |
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 |