The mathematical analysis of spatiotemporal distributions in many species exhibiting different predation mechanisms has attracted considerable attention in biology and ecology. In this article, we investigated a prey-taxis model involving hunting cooperation, which has more strong coupling structures. Utilizing energy estimates and semigroup theory, the global boundness of its classical solution was established when the hunting cooperation is weak in two dimensions. By means of Lyapunov functionals, the global asymptotically stability of the non-negative constant steady-state solution for the discussed model was established under certain assumptions on parameters. These results enrich the related researches on the prey-taxis model with Lotka-Volterra functional response, which has been studied by Jin and Wang.
Citation: Xuemin Fan, Wenjie Zhang, Lu Xu. Global dynamics of a predator-prey model with prey-taxis and hunting cooperation[J]. Electronic Research Archive, 2025, 33(3): 1610-1632. doi: 10.3934/era.2025076
[1] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[2] | Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215 |
[3] | Ailing Xiang, Liangchen Wang . Boundedness of a predator-prey model with density-dependent motilities and stage structure for the predator. Electronic Research Archive, 2022, 30(5): 1954-1972. doi: 10.3934/era.2022099 |
[4] | Jialu Tian, Ping Liu . Global dynamics of a modified Leslie-Gower predator-prey model with Beddington-DeAngelis functional response and prey-taxis. Electronic Research Archive, 2022, 30(3): 929-942. doi: 10.3934/era.2022048 |
[5] | Miao Peng, Rui Lin, Zhengdi Zhang, Lei Huang . The dynamics of a delayed predator-prey model with square root functional response and stage structure. Electronic Research Archive, 2024, 32(5): 3275-3298. doi: 10.3934/era.2024150 |
[6] | Yuan Tian, Hua Guo, Wenyu Shen, Xinrui Yan, Jie Zheng, Kaibiao Sun . Dynamic analysis and validation of a prey-predator model based on fish harvesting and discontinuous prey refuge effect in uncertain environments. Electronic Research Archive, 2025, 33(2): 973-994. doi: 10.3934/era.2025044 |
[7] | Yichao Shao, Hengguo Yu, Chenglei Jin, Jingzhe Fang, Min Zhao . Dynamics analysis of a predator-prey model with Allee effect and harvesting effort. Electronic Research Archive, 2024, 32(10): 5682-5716. doi: 10.3934/era.2024263 |
[8] | Érika Diz-Pita . Global dynamics of a predator-prey system with immigration in both species. Electronic Research Archive, 2024, 32(2): 762-778. doi: 10.3934/era.2024036 |
[9] | Xiaowen Zhang, Wufei Huang, Jiaxin Ma, Ruizhi Yang . Hopf bifurcation analysis in a delayed diffusive predator-prey system with nonlocal competition and schooling behavior. Electronic Research Archive, 2022, 30(7): 2510-2523. doi: 10.3934/era.2022128 |
[10] | San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045 |
The mathematical analysis of spatiotemporal distributions in many species exhibiting different predation mechanisms has attracted considerable attention in biology and ecology. In this article, we investigated a prey-taxis model involving hunting cooperation, which has more strong coupling structures. Utilizing energy estimates and semigroup theory, the global boundness of its classical solution was established when the hunting cooperation is weak in two dimensions. By means of Lyapunov functionals, the global asymptotically stability of the non-negative constant steady-state solution for the discussed model was established under certain assumptions on parameters. These results enrich the related researches on the prey-taxis model with Lotka-Volterra functional response, which has been studied by Jin and Wang.
In natural habitats, there exist diverse interspecific interactions, such as competition, cooperation, predation, and so on. Predation serves as a fundamental interspecific relationship, which plays a pivotal role in maintaining ecological balance by regulating the populations of predators and prey. To capture the pattern of the quality of the species and the dynamics of predation, Lotka and Volterra presented a classical ordinary differential equation (ODE) model, called the predator-prey model, in the 1920s. Subsequently, many scholars began to study various predator-prey models with different predatory mechanisms, such as direct predation [1], selective predation [2,3], cooperative predation [4,5], and so on.
The random motion of species in space is a natural movement behavior, and scholars introduce random diffusion of predators and prey into the classical predator-prey model [6,7,8] to depict the spatiotemporal distribution of species. In addition, predators seek to improve their survival prospects by making directional movements toward areas with higher density of prey. To describe this phenomenon, the following partial differential equation model, called the prey-taxis model:
{ut=d1Δu−vF(u,v)+f(u),vt=d2Δv−χ∇⋅(v∇u)+bvF(u,v)−vg(v), | (1.1) |
was proposed by Kareiva and Odell [9]. The functions u(x,t) and v(x,t) stand for the density of prey and predators, respectively. The parameters di(i=1,2)>0 mean the random diffusion coefficients. The term −χ∇⋅(v∇u) signifies the directional movements of the predator toward the prey with prey-taxis sensitivity coefficient χ>0. f(u) measures the growth rate of prey, while g(v) is the mortality rate of the predator. F(u,v) denotes the interspecific interaction, also called the functional response function, and b>0 indicates the conversion efficiency. In recent years, for model (1.1) with various functional response functions (for example, Holling type [10], ratio-dependent [11] and Beddington-DeAngelis type [12,13]), many scholars have achieved numerous results, involving topics such as the global existence, boundedness, stability, traveling waves, global bifurcation, and so on. Specifically, Jin and Wang [14] provided that the solution of (1.1) with some certain assumptions on the functions F(u,v), f(u), and g(v) is globally bounded in two dimensions. In addition, they also investigated that the prey-only steady state of (1.1) with weak predation is globally asymptotically stable, while the coexistence steady state of (1.1) with strong predation and weak prey-taxis is globally asymptotically stable.
In nature, cooperative behavior of populations is a widespread and important phenomenon in ecosystems. To ensure their survival, reproduction, and development, predators often engage in cooperative hunting, such as wolves, African wild dogs, Harris' hawks, etc. In order to investigate the impact of hunting cooperation on the density of the predator and the dynamics of the ecological community, Alves and Hilker [4] discussed an ODE model, which reads
{dudt=σu(1−uκ)−F(u,v)v,dvdt=bF(u,v)v−βv, | (1.2) |
with various functional response functions F(u,v), such as Lotka-Volterra type, Holling type II, etc. Subsequently, many scholars have conducted extensive research on the properties for the solutions of (1.2) with various functional response functions, including local existence, global boundedness, stability, global bifurcation, pattern formation. Moreover, scholars also have studied a wide range of mechanisms on (1.2), such as the Allee effect [15,16], spatial diffusion [17], time delay [18], and so on. Specifically, inspired by (1.1) and (1.2), Zhang et al. [19] studied a prey-taxis model with Holling type II hunting cooperative functional response function F(u,v)=(1+av)v1+h(1+av)u, where a>0 represents the intensity of cooperative hunting among predators and h>0 is the average handing time of the predator for the prey, which is formulated as follows:
{ut=d1Δu+σu(1−uκ)−(1+av)uv1+h(1+av)u,vt=d2Δv−∇⋅(χv∇u)+(1+av)uv1+h(1+av)u−v. | (1.3) |
They demonstrated the uniform boundedness and global existence of time-varying solutions for (1.3). Concurrently, they also analyzed the stability and prey-taxis-driven instability of positive equilibrium through linearization analysis. When χ=0, Zhang [20] incorporated predator-taxis into (1.3) and established the global existence of the classical solution for (1.3) in any spatial dimension. Additionally, he analyzed the instability induced by predator-taxis.
Hunting cooperation not only affects the population size of predators, but also has a significant impact on their spatiotemporal distribution. Therefore, considering (1.1) with the Lotka-Volterra-type hunting cooperative functional response function F(u,v)=(1+av)u (e.g., see [4]), we also introduce the random movement of species, the directional movement of predators toward prey, and the intra-specific competition within predator populations (e.g., see [14,21]) into (1.2). This derived
{ut=d1Δu+σu(1−uκ)−(1+av)uv,x∈Ω,t>0,vt=d2Δv−χ∇⋅(v∇u)+(1+av)uv−βv−γv2,x∈Ω,t>0,∂u∂ν=∂v∂ν=0,x∈∂Ω,t>0,u(x,0)=u0(x),v(x,0)=v0(x),x∈Ω, | (1.4) |
in a bounded smooth domain Ω⊂R2. The constant σ>0 is the intrinsic growth rate on prey, κ>0 stands for the carrying capacity of prey, β>0 means the mortality rate of predators, and γ>0 denotes the mortality rate caused by intra-specific competition. The initial date satisfies
u0∈W1,∞(Ω) and v0∈C0(¯Ω) with u0,v0≥0. | (1.5) |
In what follows, without confusion, we shall abbreviate ∫Ωfdx as ∫Ωf for simplicity.
Main ideas and results: It is straightforward to achieve that ‖u‖L∞(Ω) is bounded by utilizing the comparison principle on parabolic equations. We remark that from a mathematical perspective, the analysis of the global boundedness of solutions for the Lotka-Volterra-type functional response function is more difficult than that for the Holling type II functional response function derived from (1.3). After all, the holling type II functional response function allows for the estimates |(1+av)u1+h(1+av)u|≤1h due to a priori estimates for u and v. However, the Lotka-Volterra-type functional response function does not possess such a directly useful property for establishing the global boundedness of solutions. Therefore, to derive the L∞(Ω) estimate for ∇u in two-dimensional space, we need an a priori ‖v‖L6(Ω) estimate. First, we construct an energy function ∫Ωv2+∫Ω|∇u|2, which can be utilized to prove the boundness of ‖∇u‖L2(Ω) and ‖v‖L2(Ω) when a is suitably small and also establish the boundedness of ∫t+τt∫Ω|Δu|2 and ∫t+τt∫Ωv3 for some appropriately small τ∈(0,1]. Based on the estimate of ‖v‖L3(Ω), we construct the energy function ∫Ω|∇u|4+∫Ωv4 and demonstrate that v and ∇u are bounded in L4(Ω), which can help to get the a priori estimate of ‖v‖L6(Ω). So far, the estimates of ‖∇u‖L∞(Ω) and ‖v‖L∞(Ω) are proved by Neumann heat semigroups. Zhang et al. primarily discussed the stability of the coexistence equilibrium for (1.3) in [19]. In this paper, by constructing suitable Lyapunov functionals, we demonstrate the long-time behavior of the prey-only and coexistence steady state of (1.4) when a falls within a specific range. These results mean that the weak hunting cooperation of predators can avoid population overcrowding, enrich the diversity of biological populations, and enhance ecological balance.
The first main result is as follows.
Theorem 1.1. Let Ω⊂R2 be a smooth bounded domain. Then (u0,v0) satisfies (1.5), if
a<γK1(weak hunting cooperation), | (1.6) |
where K1:=max{κ,‖u0‖L∞(Ω)}, and then (1.4) possesses a positive global classical solution
(u,v)∈[C0(ˉΩ×[0,∞))∩C2,1(Ω×(0,∞))]2. |
Furthermore, the solution satisfies
‖u(⋅,t)‖W1,∞(Ω)+‖v(⋅,t)‖L∞(Ω)≤K, |
where constant K>0 does not depend on t. In particular, we also have 0<u≤K1.
Our next result aims to present the impact of the cooperative hunting on the predator-prey model for its dynamic behavior. By a simple calculation, the constant equilibrium point (us,vs) of (1.4) satisfies
{us(σ−σusκ−vs−av2s)=0,vs(us+ausvs−β−γvs)=0, | (1.7) |
which admits three possible homogeneous equilibria:
∙ Extinction steady state (0,0).
∙ Prey-only steady state (κ,0).
∙ Coexistence steady state (u∗,v∗).
Set
h(x):=a2κx3+2aκx2+(σγ+κ−aσκ)x+σ(β−κ), |
with x∈(0,+∞), and its derivative functions are
h′(x)=3a2κx2+4aκx+σγ+κ−aσκ,h″(x)=6a2κx+4aκ. |
It is easy to verify that h″(x)>0, which implies that h′(x) is strictly monotonically increasing on (0,∞).
On the other hand, from the equations presented in (1.7), we can calculate that the equilibrium point u∗ satisfies
u∗=β+γv∗1+av∗, |
where v∗ is the positive root of the equation h(x)=0. In order to compute the value of v∗, we undertake an analysis encompassing the following three cases:
Case 1: κ>β. It follows from h(0)<0 and the Descartes' rule of signs[22] that h(x)=0 has a unique root v∗∈(0,σ) and then u∗=β+γv∗1+av∗∈(β,β+γσ).
Case 2: κ=β. It holds that h(0)=0. When a≤σγ+κσκ, h′(x)>0. Thus h(x)=0 does not have a positive root. If a>σγ+κσκ, this deduces that h′(x)<0 on x∈(0,v1) and h′(x)>0 on x∈(v1,+∞), where
v1=√κ(κ+3aσκ−3σγ)−2κ3aκ. |
Therefore, h(x)=0 has a unique root v2=√σκ(aκ−γ)−κaκ>0.
Case 3: κ<β. Note that h(0)>0. If a≤σγ+κσκ, h(x)=0 has no positive root. While a>σγ+κσκ, h(x)=0 admits three statuses: no positive root as h(v1)>0, one positive root v1 as h(v1)=0, and two positive roots v3,v4 as h(v1)<0, where v3∈(0,v1) and v4∈(v1,σ).
All in all, the coexistence steady state of (1.4) satisfies
(u∗,v∗)={(u∗,v∗), if κ>β,(β+γv21+av2,v2), if κ=β and a>σγ+κσκ,(β+γv11+av1,v1), if κ<β,a>σγ+κσκ, and h(v1)=0,(β+γv31+av3,v3) and (β+γv41+av4,v4), if κ<β,a>σγ+κσκ, and h(v1)<0. |
Theorem 1.2. Suppose that the assumptions of Theorem 1.1 hold. Then:
1) Let κ≤β. If the model parameters satisfy
a<γK1(weak hunting cooperation), |
then for all t>0, the classical solution (u,v) in (1.4) converges to (κ,0) in an exponential manner, as described below:
‖u−κ‖L∞(Ω)+‖v‖L∞(Ω)≤Ce−λt, |
where constants C>0 and λ>0 are independent of t.
2) Let κ>β. If the model parameters satisfy
a<min{γK1,2√(β+σγ)2+σκγ−2(β+σγ)σκ}(weaker hunting cooperation) |
and
χ2<4d1d2u∗K21v∗, |
where u∗ and v∗ are independent of χ, then for all t>0, the classical solution (u,v) converges to (u∗,v∗) in an exponential manner, as described below:
‖u−u∗‖L∞(Ω)+‖v−v∗‖L∞(Ω)≤Ce−λt, |
where constants C>0 and λ>0 are independent of t.
Remark 1.1. Without the hunting cooperation (i.e., a=0), our results of Theorem 1.2 are consistent with the results of Proposition 1.6 in [14]. In fact, Theorem 1.2 shows that the hunting cooperation mechanism does not change the stability of (1.4) solutions when a is appropriately small.
Remark 1.2. When κ≤β, we are unable to verify the long-time behavior of the coexistence steady state (u∗,v∗) on account of the range of the intensity of cooperative hunting among predators conflicts with the condition of Theorem 1.1. However, for one-dimensional cases, the long-time behavior of the coexistence steady state is an open problem when κ≤β and a is appropriately large.
First, we provide the local existence of solutions for (1.4).
Lemma 2.1. Provided that Ω⊂R2 be a smooth bounded domain, and (u0,v0) satisfies (1.5), if the condition (1.6) holds, there exists Tmax∈(0,∞] ensuring that (1.4) possesses a classical solution
(u,v)∈[C0(ˉΩ×[0,Tmax))∩C2,1(Ω×(0,Tmax))]2 |
satisfying u,v>0. Furthermore, if Tmax<∞, then
lim supt↗Tmax{‖u(⋅,t)‖W1,∞(Ω)+‖v(⋅,t)‖L∞(Ω)}=∞. |
The conclusions of Lemma 2.1 are established from Amann's theorem [23,24].
Lemma 2.2. Provided that the assumptions of Lemma 2.1 hold, then for all t∈(0,Tmax), we have
‖u‖L∞(Ω)≤K1, | (2.1) |
where K1:=max{κ,‖u0‖L∞(Ω)}.
Proof. This lemma has been proven by using an approach similar to [25].
The following lemma gives a fundamental inequality.
Lemma 2.3. [26] Assume that Ω is a smooth bounded domain, and let g∈C2(ˉΩ) satisfy ∂g∂ν=0 on ∂Ω. Then there exists an upper bound l=l(Ω) of the curvatures of ∂Ω guaranteeing that
∂|∇g|2∂ν≤l|∇g|2. |
In order to prove Lemma 3.3, we need a lemma.
Lemma 2.4. [27] Suppose that T>0,τ∈(0,T),m1>0, and m2>0. Provided that φ:[0,T)→[0,∞) is absolutely continuous, and satisfies
φ′(t)+φ1+θ(t)≤ϕ(t)φ(t)+ψ(t),t∈R, |
where the constant θ>0, the functions ϕ(t),ψ(t)∈L1loc([0,T)) are nonnegative and
∫tt−τϕ(s)ds≤m1,∫tt−τψ(s)ds≤m2,t∈[τ,T). |
Then we can obtain
φ(t)≤φ(t0)e∫tt0ϕ(s)ds+∫tt0ψ(τ)e∫tτϕ(s)dsdτ |
and
suptφ(t)≤θ(2A1+θ)1+θθ+2Bt>t0, |
where
A=τ−11+θ(1+m1)11+θe2m1,B=τ−11+θm11+θ2e2m1+2m2e2m1+φ(0)em1. |
This section is devoted to establishing Theorem 1.1.
Lemma 3.1. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then for all t∈(0,Tmax), the classical solution (u,v) of (1.4) satisfies
‖u‖L1(Ω)+‖v‖L1(Ω)≤K2, | (3.1) |
where constant K2>0 does not depend on t.
Proof. (1.4) implies
ddt(∫Ωu+∫Ωv)+β(∫Ωu+∫Ωv)=(σ+β)∫Ωu−σκ∫Ωu2−γ∫Ωv2≤(σ+β)∫Ωu−σκ∫Ωu2. | (3.2) |
Then, using Young's inequality yields
(σ+β)∫Ωu≤σ2κ∫Ωu2+κ(σ+β)2|Ω|2σ. | (3.3) |
Combining (3.3) and (3.2), we deduce
ddt(∫Ωu+∫Ωv)+β(∫Ωu+∫Ωv)≤κ(σ+β)2|Ω|2σ, |
which implies (3.1) by ODE comparison.
Then we prove that ‖∇u‖L2(Ω) and ‖v‖L2(Ω) are bounded.
Lemma 3.2. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then for all t∈(0,Tmax), the classical solution (u,v) of (1.4) satisfies
‖∇u‖L2(Ω)+‖v‖L2(Ω)≤K3, | (3.4) |
and for all t∈(0,Tmax−τ), it holds that
∫t+τt∫Ω|Δu|2+∫t+τt∫Ωv3≤K4,with0<τ<min{1,12Tmax}, | (3.5) |
where constants K3>0 and K4>0 do not depend on t.
Proof. Intergating the sum of −Δuu times the first equation of (1.4) by parts, we have
−∫ΩutuΔu+d1∫Ω|Δu|2u=σ∫Ω(uκ−1)Δu+∫Ω(1+av)vΔu=−σκ∫Ω|∇u|2+∫ΩvΔu−2a∫Ωv∇u⋅∇v. | (3.6) |
Note that
−∫ΩutuΔu=∫Ω∇u⋅(∇uu)t=ddt∫Ω|∇u|2u−12∫Ω(|∇u|2)tu=12ddt∫Ω|∇u|2u−12∫Ω|∇u|2u2ut, |
which substituted into (3.6) gives us
12ddt∫Ω|∇u|2u+d1∫Ω|Δu|2u+σκ∫Ω|∇u|2=12∫Ω|∇u|2u2ut+∫ΩvΔu−2a∫Ωv∇u⋅∇v=d12∫Ω|∇u|2u2Δu+σ2∫Ω|∇u|2(1−uκ)u−12∫Ω|∇u|2v(1+av)u+∫ΩvΔu−2a∫Ωv∇u⋅∇v≤d12∫Ω|∇u|2u2Δu+σ2∫Ω|∇u|2u+∫ΩvΔu−2a∫Ωv∇u⋅∇v. | (3.7) |
Using ∇u⋅∇Δu=12Δ|∇u|2−|D2u|2, we conclude
d1∫Ω|∇u|2u2Δu=d1∫Ω|Δu|2u−d1∫Ω|D2u|2u+d12∫ΩΔ|∇u|2u=d1∫Ω|Δu|2u−d1∫Ω|D2u|2u+d12∫Ω∇|∇u|2⋅∇uu2+d12∫∂Ω∂|∇u|2∂ν1uds=d1∫Ω|Δu|2u−d1∫Ω|D2u|2u+d1∫Ω|∇u|4u3−d12∫Ω|∇u|2u2Δu+d12∫∂Ω∂|∇u|2∂ν1uds. | (3.8) |
Note that
∫Ωu|D2lnu|2=∫Ω|D2u|2u−2∫Ω(D2u⋅∇u)⋅∇uu2+∫Ω|∇u|4u3=∫Ω|D2u|2u+∫Ω|∇u|4u3+(∫Ω|∇u|2u2Δu−2∫Ω|∇u|4u3)=∫Ω|D2u|2u−∫Ω|∇u|4u3+∫Ω|∇u|2u2Δu. | (3.9) |
Combining (3.8) and (3.9) yields
d12∫Ω|∇u|2u2Δu=d12∫∂Ω∂|∇u|2∂ν1uds+d1∫Ω|∇u|2u−d1∫Ωu|D2lnu|2. | (3.10) |
Bringing (3.10) into (3.7), it holds that
12ddt∫Ω|∇u|2u+d1∫Ωu|D2lnu|2+σκ∫Ω|∇u|2≤d12∫∂Ω∂|∇u|2∂ν1uds+σ2∫Ω|∇u|2u+∫ΩvΔu−2a∫Ωv∇u⋅∇v. | (3.11) |
Multiplying both sides of the second equation of (1.4) by 2avχ, we derive
aχddt∫Ωv2+2ad2χ∫Ω|∇v|2=2aχ∫Ω(1+av)uv2−2aβχ∫Ωv2−2aγχ∫Ωv3+2a∫Ωv∇u⋅∇v≤2aK1χ∫Ωv2−2a(γ−aK1)χ∫Ωv3+2a∫Ωv∇u⋅∇v. | (3.12) |
Utilizing the condition a<γK1, (3.11), and (3.12), one has
ddt(12∫Ω|∇u|2u+aχ∫Ωv2)+2ad2χ∫Ω|∇v|2+d1∫Ωu|D2lnu|2+σκ∫Ω|∇u|2+2a(γ−aK1)χ∫Ωv3≤d12∫∂Ω∂|∇u|2∂ν1uds+σ2∫Ω|∇u|2u+∫ΩvΔu+2aK1χ∫Ωv2. | (3.13) |
In the light of the inequality ∫Ωu|D2lnu|2≥C1(∫Ω|D2u|2u+∫Ω|∇u|4u3) with the constant C1>0, plugged into (3.13) gives
ddt(12∫Ω|∇u|2u+aχ∫Ωv2)+2ad2χ∫Ω|∇v|2+σκ∫Ω|∇u|2+d1C1(∫Ω|D2u|2u+∫Ω|∇u|4u3)+2a(γ−aK1)χ∫Ωv3≤d12∫∂Ω∂|∇u|2∂ν1uds+σ2∫Ω|∇u|2u+∫ΩvΔu+2aK1χ∫Ωv2. | (3.14) |
Combining the trace inequality [28], the Cauchy-Schwarz inequality, and Lemma 2.3, we have
d12∫∂Ω∂|∇u|2∂ν1uds≤ld1∫∂Ω|∇u|2uds≤d1C12∫Ω(|D2u|2u+|∇u|4u)+C2∫Ω|∇u|2u, | (3.15) |
where constant C2>0. According to (2.1) and Hölder's inequality, then
(12+σ+2C22)∫Ω|∇u|2u≤(σ+1+2C22)(∫Ω|∇u|4u3)12(∫Ωu)12≤d1C14∫Ω|∇u|4u3+C3, | (3.16) |
where constant C3>0. Ustilizing |Δu|≤√2|D2u|, (2.1), and Young's inequality, we achieve
∫ΩvΔu≤d1C18∫Ω|Δu|2u+2d1C1∫Ωuv2≤d1C14∫Ω|D2u|2u+2K1d1C1∫Ωv2. | (3.17) |
We infer from the Gagliardo-Nirenberg inequality and (3.1) that
(aχ+2K1d1C1+2aK1χ)∫Ωv2=C4‖v‖2L2(Ω)≤C4C5(‖∇v‖L2(Ω)‖v‖L1(Ω)+‖v‖2L1(Ω))≤C4C5K2‖∇v‖L2(Ω)+C4C5K22≤ad2χ∫Ω|∇v|2+C6, | (3.18) |
where constants Ci(i=4,5,6)>0. Substituting (3.14)–(3.18) and |Δu|≤√2|D2u| yields
ddt(12∫Ω|∇u|2u+aχ∫Ωv2)+(12∫Ω|∇u|2u+aχ∫Ωv2)+ad2χ∫Ω|∇v|2+d1C18(∫Ω|Δu|2u+∫Ω|∇u|4u3)+σκ∫Ω|∇u|2+2a(β−aK1)χ∫Ωv3≤C3+C6, | (3.19) |
which gives (3.4) by the fact 0<u≤K1 and ODE comparison. Then integrating (3.19) over (t,t+τ) gives (3.5).
Next, we get the a priori estimate of ‖v‖L3(Ω).
Lemma 3.3. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then, the classical solution (u,v) of (1.4) satisfies
‖v‖L3(Ω)≤K5, | (3.20) |
where constant K5>0 does not depend on t.
Proof. By direct computations, we have
13ddt∫Ωv3+β∫Ωv3≤−d2∫Ωv|∇v|2+χ2d2∫Ωv3|∇u|2+K1∫Ωv3−(γ−aK1)∫Ωv4≤−d2∫Ωv|∇v|2+χ2d2(∫Ωv6)12(∫Ω|∇u|4)12+K1∫Ωv3. | (3.21) |
Applying the Gagliardo-Nirenberg inequality as n=2 yields
‖v32‖2L4(Ω)≤C1(‖∇v32‖L2(Ω)‖v32‖L2(Ω)+‖v32‖2L2(Ω)) | (3.22) |
and
‖∇u‖2L4(Ω)≤C2(‖Δu‖L2(Ω)‖∇u‖L2(Ω)+‖∇u‖2L2(Ω))≤C2(K3‖Δu‖L2(Ω)+K23) | (3.23) |
where constants C1>0 and C2>0. Thanks to (3.22), (3.23), and Young's inequality, one gets
χ2d2(∫Ωv6)12(∫Ω|∇u|4)12≤χ2C1C2d2(‖∇v32‖L2(Ω)‖v32‖L2(Ω)+‖v32‖2L2(Ω))(K3‖Δu‖L2(Ω)+K23)≤χ2C1C2K3d2‖∇v32‖L2(Ω)‖v32‖L2(Ω)‖Δu‖L2(Ω)+χ2C1C2K23d2‖v32‖2L2(Ω)+χ2C1C2K23d2‖∇v32‖L2(Ω)‖v32‖L2(Ω)+χ2C1C2K3d2‖v32‖2L2(Ω)‖Δu‖L2(Ω)≤d2∫Ωv|∇v|2+C3‖Δu‖2L2(Ω)‖v‖3L3(Ω)+C4‖v‖3L3(Ω),´ | (3.24) |
where constants C3>0 and C4>0. Substituting (3.24) into (3.21) results in
ddt‖v‖3L3(Ω)≤C3‖Δu‖2L2(Ω)‖v‖3L3(Ω)+(C4+K1)‖v‖3L3(Ω), |
which gives (3.20) by (3.5) and Lemma 2.4.
Subsequently, we achieve the bounds ‖v‖L4(Ω) and ‖∇u‖L4(Ω).
Lemma 3.4. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then the classical solution (u,v) of (1.4) satisfies
‖v‖L4(Ω)+‖∇u‖L4(Ω)≤K6, | (3.25) |
where constant K6>0 does not depend on t.
Proof. Multiplying both side of the second equation of (1.4) by v3, we deduce
14ddt∫Ωv4+∫Ωv4≤−3d2∫Ωv2|∇v|2+3χ∫Ωv3∇u⋅∇v+(K1+1)∫Ωv4−(γ−aK1)∫Ωv5≤−2d2∫Ωv2|∇v|2+9χ24d2∫Ωv4|∇u|2+(K1+1)∫Ωv4. | (3.26) |
Integrating the first equation of (1.4), while taking into account the fact that 0<u≤K1, it follows that
14ddt∫Ω|∇u|4+d12∫Ω|∇|∇u|2|2+d1∫Ω|∇u|2|D2u|2=d12∫∂Ω|∇u|2∂|∇u|2∂νds+σ∫Ω|∇u|4(1−2κu)−∫Ω|∇u|2∇u⋅∇(uv)−a∫Ω|∇u|2∇u⋅∇(uv2)=d12∫∂Ω|∇u|2∂|∇u|2∂νds+σ∫Ω|∇u|4(1−2κu)+∫Ωuv|∇u|2Δu+a∫Ωuv∇(|∇u|2)∇u+∫Ωuv2|∇u|2Δu+a∫Ωuv2∇(|∇u|2)∇u≤d12∫∂Ω|∇u|2∂|∇u|2∂νds+K1∫Ωv(|∇u|2|Δu|+|∇|∇u|2||∇u|)+aK1∫Ωv2(|∇u|2|Δu|+|∇|∇u|2||∇u|)+σ∫Ω|∇u|4. | (3.27) |
Thanks to the trace inequality and Lemma 2.3, we know
d12∫∂Ω|∇u|2∂|∇u|2∂νds≤ld12‖|∇u|2‖2L2(∂Ω)≤d18∫Ω|∇|∇u|2|2+C1∫Ω|∇u|4. | (3.28) |
Utilizing |Δu|≤√2|D2u| and ∇|∇u|2=2D2u⋅∇u gives rise to
K1∫Ωv(|∇u|2|Δu|+|∇|∇u|2||∇u|)≤√2K1∫Ωv|∇u|2|D2u|+2K1∫Ωv|∇u|2|D2u|=(√2+2)K1∫Ωv|∇u|2|D2u|≤d14∫Ω|∇u|2|D2u|2+K1∫Ωv4+(√2+2)4K314d21∫Ω|∇u|4 | (3.29) |
and
aK1∫Ωv2(|∇u|2|Δu|+|∇|∇u|2||∇u|)≤√2aK1∫Ωv2|∇u|2|D2u|+2aK1∫Ωv|∇u|2|D2u|=(√2+2)aK1∫Ωv2|∇u|2|D2u|≤d14∫Ω|∇u|2|D2u|2+(√2+2)2a2K21d1∫Ωv4|∇u|2. | (3.30) |
The Gagliardo-Nirenberg inequality implies
(σ+1+C1+(√2+2)4K314d21)∫Ω|∇u|4=(σ+1+C1+(√2+2)4K314d21)‖|∇u|2‖2L2(Ω)≤C2‖∇|∇u|2‖L2(Ω)‖|∇u|2‖L1(Ω)+C2‖|∇u|2‖2L1(Ω)≤C2K23‖∇|∇u|2‖L2(Ω)+C2K43≤d18∫Ω|∇|∇u|2|2+C3, | (3.31) |
where constants C2>0 and C3>0. Adding (3.27)–(3.31) and (3.26), we conclude
14ddt(∫Ωv4+∫Ω|∇u|4)+∫Ωv4+∫Ω|∇u|4+2d2∫Ωv2|∇v|2+d14∫Ω|∇|∇u|2|2+d12∫Ω|∇u|2|D2u|2≤(9χ24d2+(√2+2)2a2K21d1)∫Ωv4|∇u|2+(2K1+1)∫Ωv4+C3≤C4(∫Ωv6)23(∫Ω|∇u|6)13+(2K1+1)∫Ωv4+C3, | (3.32) |
where constant C4>0. Applying (3.4) and the Gagliardo-Nirenberg inequality yields
(2K1+1)∫Ωv4=(2K1+1)‖v2‖L2(Ω)≤(2K1+1)C5(‖∇v2‖L2(Ω)‖v2‖L1(Ω)+‖v2‖2L1(Ω))≤(2K1+1)K23C5‖∇v2‖L2(Ω)+(2K1+1)K43C5≤d2∫Ωv2|∇v|2+C6, | (3.33) |
where constants C5>0 and C6>0. In addition,
‖∇u‖2L6(Ω)=‖|∇u|2‖L3(Ω)≤C7‖∇|∇u|2‖23L2(Ω)‖|∇u|2‖13L1(Ω)+C7‖|∇u|2‖L1(Ω)≤C7K233‖∇|∇u|2‖23L2(Ω)+C7K23, | (3.34) |
where constant C7>0. Applying Young's inequality along with (3.34), one has
C4‖v2‖2L3(Ω)‖∇u‖2L6(Ω)≤C4C7K233‖v2‖2L3(Ω)‖∇|∇u|2‖23L2(Ω)+C4C7K23‖v2‖2L3(Ω)≤d14‖∇|∇u|2‖2L2(Ω)+C8‖v2‖3L3(Ω)+C9, | (3.35) |
where constants C8>0 and C9>0. Utilizing the Gagliardo-Nirenberg inequality and Lemma 3.3 yields
C8‖v2‖3L3(Ω)≤C8C10(‖∇v2‖32L2(Ω)‖v2‖32L32(Ω)+‖v2‖2L32(Ω))≤C8C10K35‖∇v2‖32L2(Ω)+C8C10K45≤d2∫Ωv2|∇v|2+C11, | (3.36) |
where constants C10>0 and C11>0. Plugging (3.33), (3.35), and (3.36) into (3.32), there exists a constant C12=C3+C6+C9+C11 such that
14ddt(∫Ωv4+∫Ω|∇u|4)+∫Ωv4+∫Ω|∇u|4≤C12, |
which in conjunction with ODE comparison results in (3.25).
Lemma 3.5. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then for all t∈(0,Tmax), we can get
‖v‖L6(Ω)≤K7, | (3.37) |
where constant K7>0 does not depend on t.
Proof. We can directly compute
16ddt∫Ωv6+β∫Ωv6+5d2∫Ωv4|∇v|2=5χ∫Ωv5∇u⋅∇v+∫Ωuv6+a∫Ωuv7−γ∫Ωv7≤d2∫Ωv4|∇v|2+25χ24d2∫Ωv6|∇u|2+K1∫Ωv6−(γ−aK1)∫Ωv7≤d2∫Ωv4|∇v|2+25χ24d2(∫Ωv12)12(∫Ω|∇u|4)12+K1∫Ωv6≤d2∫Ωv4|∇v|2+25χ2K264d2(∫Ωv12)12+K1∫Ωv6. | (3.38) |
Utilizing the Gagliardo-Nirenberg inequality, it follows that
25χ2K264d2(∫Ωv12)12=25χ2K264d2‖v3‖2L4(Ω)≤25χ2K26C14d2(‖∇v3‖32L2(Ω)‖v3‖12L1(Ω)+‖v3‖2L1(Ω))≤25χ2C1K26K3254d2‖∇v3‖32L2(Ω)+25χ2C1K26K654d2≤d2∫Ωv4|∇v|2+C2, | (3.39) |
where constants C1>0 and C2>0. Then
K1∫Ωv6=K1‖v3‖2L2(Ω)≤K1C3(‖∇v3‖L2(Ω)‖v3‖L1(Ω)+‖v3‖2L1(Ω))≤C3K1K35‖∇v3‖L2(Ω)+C3K1K65≤d2∫Ωv4|∇v|2+C4, | (3.40) |
where constants C3>0 and C4>0. Plugging (3.39) and (3.40) into (3.38), we conclude
16ddt∫Ωv6+β∫Ωv6≤C2+C4, |
which results in (3.37) with ODE comparison.
We are currently capable of deducing the bounded property of ‖∇u‖L∞(Ω) and ‖v‖L∞(Ω) in the case where the dimension n=2.
Lemma 3.6. Provided that the assumptions of Lemma 2.1 are fulfilled, if the condition (1.6) holds, then for all t∈(0,Tmax), we have
‖∇u‖L∞(Ω)≤K8 | (3.41) |
and
‖v‖L∞(Ω)≤K9, | (3.42) |
where constants K8>0 and K9>0 do not depend on t.
Proof. The variation-of-constants formula implies
u(⋅,t)=ed1tΔu0+∫t0ed1(t−s)Δ[σu(1−uκ)−(1+av)uv]ds, |
and hence
∇u(⋅,t)=∇ed1tΔu0+∫t0∇ed1(t−s)Δ[σu(1−uκ)−(1+av)uv]ds. |
Then by (3.1), (3.20), and (3.37), it holds that
‖σu(1−uκ)−(1+av)u)‖L3(Ω)≤‖σu−σκu2‖L3(Ω)+‖uv‖L3(Ω)+a‖uv2‖L3(Ω)≤σK1(1+K1κ)|Ω|13+K1K5+aK1K27. | (3.43) |
Applying the Neumann heat semigroup [29] and (3.43), there exist constants γ1>0 and λ1>0 ensuring that
‖∇u‖L∞(Ω)≤‖∇ed1tΔu0‖L∞(Ω)+∫t0‖∇ed1(t−s)Δ[σu(1−uκ)−(1+av)uv]‖L∞(Ω)ds≤2γ1e−d1λ1t‖u0‖L∞(Ω)+γ1∫t0(1+(t−s)−12−13)e−d1λ1t‖σu(1−uκ)−(1+av)uv‖L3(Ω)ds≤2γ1‖u0‖L∞(Ω)+γ1[σK1(1+K1κ)|Ω|13+K1K5+aK1K27]∫∞0(1+(t−s)−56)e−d1λ1tds≤2γ1‖u0‖L∞(Ω)+γ1d1λ1[σK1(1+K1κ)|Ω|13+K1K5+aK1K27](1+(d1λ1)56Γ(16)), |
which gives (3.41) and gamma function Γ(z)=∫∞0tz−1e−tdt. We use (3.20) and (3.41) to obtain
‖v∇u‖L3(Ω)≤‖∇u‖L∞(Ω)‖v‖L3(Ω)≤K5K8 | (3.44) |
and combining with (3.1), (3.20), and (3.37), we conclude
‖(1+av)uv‖L3(Ω)≤‖u‖L∞(Ω)(‖v‖L3(Ω)+a‖v2‖L3(Ω))≤K1(K5+aK27). | (3.45) |
Applying the Neumann heat semigroup [29], (3.44), and(3.45), there exist constants γ2>0, γ3>0, and λ1>0 ensuring that
‖v‖L∞(Ω)≤‖et(d2Δ−1)v0‖L∞(Ω)+χ∫t0‖e(t−s)(d2Δ−1)∇⋅(v∇u)‖L∞(Ω)ds+∫t0‖e(t−s)(d2Δ−1)(1+av)uv‖L∞(Ω)ds≤‖v0‖L∞(Ω)+γ2χ∫t0(1+(t−s)−56)e−(λ1d2+1)(t−s)‖v∇u‖L3(Ω)ds+γ3∫t0(1+(t−s)−13e−(t−s))‖(1+av)uv‖L3(Ω)ds≤‖v0‖L∞(Ω)+γ2χK5K8∫∞0(1+(t−s)−56)e−(t−s)ds+γ3K1(K5+aK27)∫∞0(1+(t−s)−13)e−(t−s)ds≤‖v0‖L∞(Ω)+γ2χK5K8(1+Γ(16))+γ3K1(K5+aK27)(1+Γ(23)), |
which gives (3.42).
Proof of Theorem 1.1. Combining Lemma 2.2 and Lemma 3.6, then, there exists a positive constant K which guarantees that
‖u(⋅,t)‖W1,∞(Ω)+‖v(⋅,t)‖L∞(Ω)≤K, |
which proves Theorem 1.1 with the extension criterion outlined in Lemma 2.1.
Based on Theorem 1.1, this part aims to establish the convergence property of the solution. Before this, we introduce a lemma as follows.
Lemma 4.1. For all given θ∈(0,1), we have
‖u‖C2+θ,1+θ2(ˉΩ×[t,t+1])+‖v‖C2+θ,1+θ2(ˉΩ×[t,t+1])≤C,t≥1 |
with a constant C>0.
Proof. The derived conclusion stems directly from the regularity properties of parabolic equations as outlined in [30].
In this part, we will demonstrate that the solution (u,v) converges to (κ,0) under certain conditions. To achieve this aim, we introduce the Lyapunov functional, denoted as
F1(t):=∫Ω(u−κ−κlnuκ)+∫Ωv,t>0. |
Lemma 4.2. Under the assumed conditions of Theorem 1.1, if the parameter satisfies κ≤β and a<γK1, this ensures that
‖u−κ‖L∞(Ω)+‖v‖L∞(Ω)≤K10e−δ1t,t>0, |
where constants K10>0 and δ1>0.
Proof. From (1.4), we have
ddt∫Ω(u−κ−κlnuκ)=∫Ω(1−κu)(d1Δu+σu(1−uκ)−(1+av)uv)=−d1κ∫Ω|∇u|2u2−σκ∫Ω(u−κ)2+κ∫Ωv+aκ∫Ωv2−∫Ωuv−a∫Ωuv2 | (4.1) |
and
ddt∫Ωv=∫Ω(d2Δv−χ∇⋅(v∇u)+(1+av)uv−βv−γv2)=∫Ωuv+a∫Ωuv2−β∫Ωv−γ∫Ωv2. | (4.2) |
Adding (4.1) and (4.2) results in
ddtF1(t)≤−d1κ∫Ω|∇u|2u2−σκ∫Ω(u−κ)2−(β−κ)∫Ωv−(γ−aκ)∫Ωv2. | (4.3) |
Then, due to κ≤β and a<γK1, one can choose a constant c1>0 such that
ddtF1(t)≤−c1(∫Ω(u−κ)2+∫Ωv+∫Ωv2),t>0, |
which provides
∫+∞1∫Ω(u−κ)2+∫+∞1∫Ωv2≤c2, | (4.4) |
where constant c2>0. Using (4.4) and the uniform continuity of u and v due to Lemma 4.1 yields
∫Ω(u−κ)2+∫Ωv2→0, as t→+∞. |
Utilizing the Gagliardo-Nirenberg inequality, for all t>1, we derive
‖u−κ‖L∞(Ω)≤c3‖u−κ‖2N+2W1,∞(Ω)‖u−κ‖2N+2L2(Ω) | (4.5) |
and
‖v‖L∞(Ω)≤c4‖v‖2N+2W1,∞(Ω)‖v‖2N+2L2(Ω), | (4.6) |
where constants c3>0 and c4>0, which proves the claim with Lemma 4.1 and (4.4). Furthermore, applying L'Hoptial's rule, it holds that
lims→s0s−s0−s0lnss0(s−s0)2=lims→s01−s0s2(s−s0)=lims→s012s=12s0,s0>0. |
There exists a constant ε>0 ensuring that
14s0(s−s0)2≤s−s0−s0lnss0≤1s0(s−s0)2 for all |s−s0|≤ε. | (4.7) |
By (4.5) and (4.6), there exists t0>0 ensuring that
‖u−κ‖L∞(Ω)+‖v‖L∞(Ω)≤ε,t≥t0. |
Therefore, by (4.7), we get
14κ∫Ω(u−κ)2≤∫Ω(u−κ−κlnuκ)≤1κ∫Ω(u−κ)2,t≥t0, |
which gives
F1(t)≤1c5(∫Ω(u−κ)2+∫Ωv2),t≥t0, | (4.8) |
where constant c5>0. Plugging (4.8) into (4.3), it follows that
ddtF1(t)≤−(∫Ω(u−κ)2+∫Ωv2)≤−c5F1(t),t≥t0. |
Therefore, we obtain
F1(t)≤e−c5(t−t0)F1(t0)+c5(t−t0)e−c5t≤c6e−δ1t,t≥t0 |
with some constants c6>0 and δ1>0.
In this part, we will demonstrate that the solution (u,v) converges to (u∗,v∗) under certain conditions. To achieve this aim, we define the Lyapunov functional
F2(t):=∫Ω(u−u∗−u∗lnuu∗)+∫Ω(v−v∗−v∗lnvv∗),t>0, |
where u∗∈(β,β+γσ) and v∗∈(0,σ).
Lemma 4.3. Under the assumed conditions of Theorem 1.1, if the parameters satisfy κ>β and
a<min{γK1,2√(β+σγ)2+σκγ−2(β+σγ)σκ}, | (4.9) |
as well as
χ2<4d1d2u∗K21v∗, | (4.10) |
this ensures that
‖u−u∗‖L∞(Ω)+‖v−v∗‖L∞(Ω)≤K11e−δ2t,t>0, |
where constants K11>0 and δ2>0.
Proof. Using σ(1−u∗κ)−(1+av∗)v∗=0, it can be directly calculated that
ddt∫Ω(u−u∗−u∗lnuu∗)=∫Ω(1−u∗u)(d1Δu+σu(1−uκ)−(1+av)uv)=−d1u∗∫Ω|∇u|2u2−σκ∫Ω(u−u∗)2−∫Ω(u−u∗)(v−v∗)−a∫Ω(u−u∗)(v2−v2∗). | (4.11) |
Utilizing (1+av∗)u∗−β−γv∗=0, we have
ddt∫Ω(v−v∗−v∗lnvv∗)=∫Ω(1−v∗v)(d2Δv−χ∇⋅(v∇u)+(1+av)uv−βv−γv2)=−d2v∗∫Ω|∇v|2v2+χv∗∫Ω∇u⋅∇vv−γ∫Ω(v−v∗)2+∫Ω(u−u∗)(v−v∗)+a∫Ω(v−v∗)(uv−u∗v∗). | (4.12) |
Combining (4.11) and (4.12), it follows that
ddtF2(t)=−d1u∗∫Ω|∇u|2u2−d2v∗∫Ω|∇v|2v2+χv∗∫Ω∇u⋅∇vv−σκ∫Ω(u−u∗)2−γ∫Ω(v−v∗)2−a∫Ω(v−v∗)(uv∗−u∗v)=−d1u∗∫Ω|∇u|2u2−d2v∗∫Ω|∇v|2v2+χv∗∫Ω∇u⋅∇vv−σκ∫Ω(u−u∗)2−∫Ω(γ−au∗)(v−v∗)2−av∗∫Ω(u−u∗)(v−v∗):=−XQXT−YMYT, | (4.13) |
where X=(∇uu,∇vv), Y=(u−u∗,v−v∗), and the matrices Q and M stand for
Q=(d1u∗−χv∗u2−χv∗u2d2v∗) |
and
M=(σκav∗2av∗2γ−au∗). |
If (4.9) and (4.10) hold, we check
|Q|=d1d2u∗v∗−χ2v2∗u24>d1d2u∗v∗−χ2v2∗K214>0 |
and
|M|=σκ(γ−au∗)−a2v2∗4>0, |
which means that the matrices Q and M are positive definite with K1:=max{κ,‖u0‖L∞(Ω)}. Then for all u,v, it implies that
ddtF2(t)≤−c1∫Ω(|∇u|2u2+|∇v|2v2)−c2∫Ω((u−u∗)2+(v−v∗)2), |
where constants c1>0 and c2>0. Subsequently, the rest is analogous to the reasoning employed in Lemma 4.2, and we readily demonstrate that the solution (u,v) exponentially converges to (u∗,v∗) as t→∞ in L∞(Ω).
Proof of Theorem 1.2. Combining Lemmas 4.2 and 4.3, we directly obtain Theorem 1.2.
In this paper, we have proposed a prey-taxis model with hunting cooperation mechanism and explored the effect of predator hunting cooperation mechanism on predator and prey populations. Through mathematical analysis, we have demonstrated that weak cooperative hunting can prevent the blow-up of the classical solutions for model (1.4) in two-dimensional space. On the one hand, we have received that the long time behavior of the prey-only steady state is established under the weak hunting cooperation. On the other hand, we have only obtained the long time behavior for the coexistence steady states under weaker hunting cooperation. However, for the strong hunting cooperation mechanism, whether the long time behavior of the coexistence steady state can be established remains an open problem that necessitates further research.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work is supported by the Science and Technology Project of YiLi Prefecture (Grant No. YZ2022B038), the Special Project of Yili Normal University to Improve Comprehensive Strength of Disciplines (Grant No. 22XKZY14), and the Natural Science Foundation of Xinjiang Autonomous Region (Grant No: 2022D01C335).
The authors declare that there is no conflict of interest.
[1] |
S. Creel, D. Christianson, Relationships between direct predation and risk effects, Trends Ecol. Evol., 23 (2008), 194–201. https://doi.org/10.1016/j.tree.2007.12.004 doi: 10.1016/j.tree.2007.12.004
![]() |
[2] |
R. A. Stein, Selective predation, optimal foraging, and the predator-prey interaction between fish and crayfish, Ecology, 58 (1977), 1237–1253. https://doi.org/10.2307/1935078 doi: 10.2307/1935078
![]() |
[3] |
S. Roy, P. K. Tiwari, H. Nayak, M. Martcheva, Effects of fear, refuge and hunting cooperation in a seasonally forced eco-epidemic model with selective predation, Eur. Phys. J. Plus., 137 (2022), 528. https://doi.org/10.1140/epjp/s13360-022-02751-2 doi: 10.1140/epjp/s13360-022-02751-2
![]() |
[4] |
M. T. Alves, F. M. Hilker, Hunting cooperation and Allee effects in predator, J. Theor. Biol., 419 (2017), 13–22. https://doi.org/10.1016/j.jtbi.2017.02.002 doi: 10.1016/j.jtbi.2017.02.002
![]() |
[5] |
T. Singh, S. Banerjee, Spatial aspect of hunting cooperation in predators with Holling type II functional response, J. Biol. Syst., 26 (2018), 511–531. https://doi.org/10.1142/S0218339018500237 doi: 10.1142/S0218339018500237
![]() |
[6] |
B. Mukhopadhyay, R. Bhattacharyya, Modeling the role of diffusion coefficients on Turing instability in a reaction-diffusion prey-predator system, Bull. Math. Biol., 68 (2006), 293–313. https://doi.org/10.1007/s11538-005-9007-2 doi: 10.1007/s11538-005-9007-2
![]() |
[7] | Y. Du, J. Shi, Some recent results on diffusive predator-prey models in spatially heterogeneous environment, in Nonlinear Dynamics and Evolution Equations, 48 (2006), 95–135. |
[8] |
F. Wang, R. Yang, X. Zhang, Turing patterns in a predator-prey model with double Allee effect, Math. Comput. Simul., 220 (2024), 170–191. https://doi.org/10.1016/j.matcom.2024.01.015 doi: 10.1016/j.matcom.2024.01.015
![]() |
[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. Wang, W. Wang, G. Zhang, Global bifurcation of solutions for a predator-prey model with prey-taxis, Math. Methods Appl. Sci., 38 (2015), 431–443. https://doi.org/10.1002/mma.3079 doi: 10.1002/mma.3079
![]() |
[11] |
Y. Cai, Q. Cao, Z. 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
![]() |
[12] |
N. K. Thakur, R. Gupta, R. K. Upadhyay, Complex dynamics of diffusive predator-prey system with Beddington-DeAngelis functional response: The role of prey-taxis, Asian-Eur. J. Math., 10 (2017), 1750047. https://doi.org/10.1142/S1793557117500474 doi: 10.1142/S1793557117500474
![]() |
[13] |
D. Luo, Q. Wang, Global bifurcation and pattern formation for a reaction-diffusion predator-prey model with prey-taxis and double Beddington-DeAngelis functional responses, Nonlinear Anal.: Real World Appl., 67 (2022), 103638. https://doi.org/10.1016/j.nonrwa.2022.103638 doi: 10.1016/j.nonrwa.2022.103638
![]() |
[14] |
H. Jin, Z. 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
![]() |
[15] |
S. R. Jang, W. Zhang, V. Larriva, Cooperative hunting in a predator-prey system with Allee effects in the prey, Nat. Resour. Model., 31 (2018), 12194. https://doi.org/10.1111/nrm.12194 doi: 10.1111/nrm.12194
![]() |
[16] |
D. Sen, S. Ghorai, M. Banerjee, Allee effect in prey versus hunting cooperation on predator-enhancement of stable coexistence, Int. J. Bifurcation Chaos, 29 (2019), 1950081. https://doi.org/10.1142/S0218127419500810 doi: 10.1142/S0218127419500810
![]() |
[17] |
X. Meng, L. Xiao, Hopf bifurcation and turing instability of a delayed diffusive zooplankton-phytoplankton model with hunting cooperation, In. J. Bifurcation Chaos, 34 (2024), 2450090. https://doi.org/10.1142/S0218127424500901 doi: 10.1142/S0218127424500901
![]() |
[18] |
I. Benamara, A. El Abdllaoui, J. Mikram, Impact of time delay and cooperation strategy on the stability of a predator-prey model with Holling type III functional response, Int. J. Biomath., 16 (2023), 2250089. https://doi.org/10.1142/S1793524522500899 doi: 10.1142/S1793524522500899
![]() |
[19] |
H. Zhang, S. Fu, C. Huang, Global solutions and pattern formations for a diffusive prey-predator system with hunting cooperation and prey-taxis, Discrete Contin. Dyn. Syst. Ser. B, 29 (2024), 3621–3644. https://doi.org/10.3934/dcdsb.2024017 doi: 10.3934/dcdsb.2024017
![]() |
[20] |
H. Zhang, Dynamics behavior of a predator-prey diffusion model incorporating hunting cooperation and predator-taxis, Mathematics, 12 (2024), 1474. https://doi.org/10.3390/math12101474 doi: 10.3390/math12101474
![]() |
[21] |
J. M. Lee, T. Hillen, M. A. Lewis, Pattern formation in prey-taxis systems, J. Biol. Dyn., 3 (2009), 551–573. https://doi.org/10.1080/17513750802716112 doi: 10.1080/17513750802716112
![]() |
[22] | D. J. Struik, A Source Book in Mathematics, 1200-1800, Harvard University Press, Cambridge, 1986. |
[23] |
H. Amann, Dynamic theory of quasilinear parabolic equations. II. Reaction-diffusion systems, Differ. Integr. Equations, 3 (1990), 13–75. https://doi.org/10.57262/die/1371586185 doi: 10.57262/die/1371586185
![]() |
[24] | H. Amann, Nonhomogeneous linear and quasilinear elliptic and parabolic boundary value problems, Function Spaces, in Function Spaces, Differential Operators and Nonlinear Analysis, 133 (1993), 9–126. https://doi.org/10.1007/978-3-663-11336-2_1 |
[25] |
Y. Tao, M. Winkler, Large time behavior in a forager-exploiter model with different taxis strategies for two groups in search of food, Math. Models Methods Appl. Sci., 29 (2019), 2151–2182. https://doi.org/10.1142/S021820251950043X doi: 10.1142/S021820251950043X
![]() |
[26] |
P. Souplet, N. Mizoguchi, Nondegeneracy of blow-up points for the parabolic keller-segel system, Ann. Inst. Henri Poincare C, Anal. Non Lineaire, 31 (2014), 851–875. https://doi.org/10.1016/j.anihpc.2013.07.007 doi: 10.1016/j.anihpc.2013.07.007
![]() |
[27] |
C. Jin, Global classical solution and stability to a coupled chemotaxis-fluid model with logistic source, Discrete Contin. Dyn. Syst., 38 (2018), 3547–3566. https://doi.org/10.3934/dcds.2018150 doi: 10.3934/dcds.2018150
![]() |
[28] | P. Quittner, P. Souplet, Superlinear Parabolic Problems: Blow-Up, Global Existence and Steady States, Birkhäuser, Basel, 2007. |
[29] |
M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model, J. Differ. Equations, 248 (2010), 2889–2905. https://doi.org/10.1016/j.jde.2010.02.008 doi: 10.1016/j.jde.2010.02.008
![]() |
[30] |
M. M. Porzio, V. Vespri, Holder estimates for local solutions of some doubly nonlinear degenerate parabolic equations, J. Differ. Equations, 103 (1993), 146–178. https://doi.org/10.1006/jdeq.1993.1045 doi: 10.1006/jdeq.1993.1045
![]() |