In this paper, we consider an initial-Neumann boundary value problem for a two-species chemotaxis system
{∂u∂t=Δu−χ∇⋅(u∇w)+u(a1−b1um−1+c1v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−ξ∇⋅(v∇w)+v(a2−b2vl−1−c2u), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−(uα+vβ)w, (x,t)∈Ω×(0,Tmax),
where the domain Ω⊂Rn(n≥2) is bounded and smooth, Tmax∈(0,∞], and parameters ai,bi,ci,m,l,α, β,χ,ξ>0 with m,l>1,i=1,2. In the current work, we provide a sufficient condition of global classical solvability to the above system. More precisely, for some suitable initial data, if m>max{α(n+2)2,1} and l>max{β(n+2)2,1}, then the system has a global classical solution. Compared to previous work, the existence result established here is more generalized, depending only on the nonlinear power exponents and spatial dimensions.
Citation: Chang-Jian Wang, Yuan-Hao Zang. Boundedness of solutions in a two-species chemotaxis system[J]. Electronic Research Archive, 2025, 33(5): 2862-2880. doi: 10.3934/era.2025126
[1] | 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 |
[2] | 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 |
[3] | Jiani Jin, Haokun Qi, Bing Liu . Hopf bifurcation induced by fear: A Leslie-Gower reaction-diffusion predator-prey model. Electronic Research Archive, 2024, 32(12): 6503-6534. doi: 10.3934/era.2024304 |
[4] | Xuemin Fan, Wenjie Zhang, Lu Xu . Global dynamics of a predator-prey model with prey-taxis and hunting cooperation. Electronic Research Archive, 2025, 33(3): 1610-1632. doi: 10.3934/era.2025076 |
[5] | Shuxia Pan . Asymptotic spreading in a delayed dispersal predator-prey system without comparison principle. Electronic Research Archive, 2019, 27(0): 89-99. doi: 10.3934/era.2019011 |
[6] | Yuan Tian, Yang Liu, Kaibiao Sun . Complex dynamics of a predator-prey fishery model: The impact of the Allee effect and bilateral intervention. Electronic Research Archive, 2024, 32(11): 6379-6404. doi: 10.3934/era.2024297 |
[7] | 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 |
[8] | 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 |
[9] | Pinglan Wan . Dynamic behavior of stochastic predator-prey system. Electronic Research Archive, 2023, 31(5): 2925-2939. doi: 10.3934/era.2023147 |
[10] | Ling Xue, Min Zhang, Kun Zhao, Xiaoming Zheng . Controlled dynamics of a chemotaxis model with logarithmic sensitivity by physical boundary conditions. Electronic Research Archive, 2022, 30(12): 4530-4552. doi: 10.3934/era.2022230 |
In this paper, we consider an initial-Neumann boundary value problem for a two-species chemotaxis system
{∂u∂t=Δu−χ∇⋅(u∇w)+u(a1−b1um−1+c1v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−ξ∇⋅(v∇w)+v(a2−b2vl−1−c2u), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−(uα+vβ)w, (x,t)∈Ω×(0,Tmax),
where the domain Ω⊂Rn(n≥2) is bounded and smooth, Tmax∈(0,∞], and parameters ai,bi,ci,m,l,α, β,χ,ξ>0 with m,l>1,i=1,2. In the current work, we provide a sufficient condition of global classical solvability to the above system. More precisely, for some suitable initial data, if m>max{α(n+2)2,1} and l>max{β(n+2)2,1}, then the system has a global classical solution. Compared to previous work, the existence result established here is more generalized, depending only on the nonlinear power exponents and spatial dimensions.
Chemotaxis refers to the phenomenon of directional movement of cells or organisms in response to chemical stimuli. The first system of partial differential equations with respect to chemotaxis was established by Keller and Segel [1] from a mathematical perspective. Thereafter, considering the influence of some factors (for instance, logistic terms [2,3], nonlinear diffusions [4,5,6], fluid effects [7,8], and the consumption mechanism [9]), many more complex variants of this model have been proposed. These models and related models also have many applications across various fields, such as ecological population models [10], pattern formation (see [11,12]), electrorheological fluids (see [13]), and image restoration (see [14,15,16]).
The chemotaxis-consumption system can be described as
{∂u∂t=Δu−χ∇⋅(u∇v)+f(u), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−uv, (x,t)∈Ω×(0,Tmax),∂u∂ν=∂v∂ν=0, (x,t)∈∂Ω×(0,Tmax),u(x,0)=u0(x),v(x,0)=v0(x), x∈Ω, | (1.1) |
where Tmax∈(0,∞] represents the maximum existence time of the solution, and u and v represent cell population density and oxygen concentration, respectively. In recent years, substantial theoretical results have been obtained regarding the related model [17,18,19]. For f(u)=0, if 0<χ<16(n+1)‖v0‖L∞(Ω), Tao [20] elaborated that the corresponding system is globally classically solvable by establishing the boundedness of a weighted functional. Baghaei and Khelghati [21] obtained the same results by improving the condition obtained in [20] with 0<χ<π√2(n+1)‖v0‖L∞(Ω). Fuest [22] considered a more generalized system with indirect consumption effect, ∂u∂t=Δu−∇⋅(u∇v),∂v∂t=Δv−vw,∂w∂t=−δw+u with δ>0, and gave some sufficient conditions for global classical solvability with n≤2 or ‖v0‖L∞(Ω)≤13n. For f(u)=au−bu2 with a,b>0, Lankeit and Wang [23] studied the influence of the size of parameter a on the global existence of solutions, including smooth solutions and weak solutions.
As demonstrated in the above models, the mechanism of resource consumption is a linear form of function u. However, based on the complexity of the external environment, the nonlinear dependence of resource dissipation on the cell density function u seems to be more reasonable sometimes. Recently, a nonlinear coupled chemotaxis-consumption problem [24] has been studied,
{∂u∂t=Δu−χ∇⋅(u∇v)+ξ∇⋅(u∇w)+au−bum, (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−uαv, (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−uβw, (x,t)∈Ω×(0,Tmax), | (1.2) |
where a,b,α,β,ξ,χ,m are positive constants. In [24], we provided a sufficient condition on the existence of classical solution with m>max{max{α,β}(n+2)2,1}. Chiyo et al. [25] studied system (1.2) involving volume-filling effect with α,β∈(0,1), and provided a detailed characterization on the global classical solvability. Afterwards, a more generalized chemotaxis system, also called the nonlinear indirect chemotaxis-consumption system, has been discussed, and similar results on classical solutions have been demonstrated [26].
Considering interactions between two species under the stimulation of chemical signal, we get the following system:
{∂u∂t=Δu−ξ∇⋅(u∇w)+f1(u,v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−χ∇⋅(v∇w)+f2(u,v), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−γw+αu+βv, (x,t)∈Ω×(0,Tmax), | (1.3) |
where α,β,γ,ξ,χ are positive constants, and the nonlinear functions f1,f2 are used to characterize the relationship between two species. For the case where f1,f2 represent the competition kinetics of two species formulated by f1(u,v)=μ1(1−u−a1v),f2(u,v)=μ2(1−a2u−v) with μi,ai>0,i=1,2, Bai and Winkler [27] discussed the corresponding system in Ω⊂Rn with n≤2 and obtained the global solvability in the classical sense. Additionally, for the case where 0<a1,a2<1 and μ1,μ2>C or 1≤a1<∞,0<a2<1 and μ2>C with some C>0, the long-time behavior of solutions was also studied therein. Mizukami [28] studied a quasilinear version of (1.3) and improved the hypothesis established in [27] by enlarging the ranges of μ1,μ2. Later, Mizukami [29] further obtained the improvement of conditions for the case a1,a2∈(0,1) based on [27,28]. For the higher-dimensional case with n≥2, the global existence in the smooth sense was explored in [30,31]. If f1,f2 are formulated by f1(u,v)=μ1u(1−u−a1v) and f2(u,v)=μ2v(1−v+a2u), then system (1.3) turns into a predator-prey system involving chemotaxis mechanisms. Subsequently, for n=3, the global classical solvability was established in [32].
More recently, when considering both species consuming nutrients, the following chemotaxis competition model has been investigated:
{∂u∂t=Δu−ξ1∇⋅(u∇w)+μ1u(1−u−a1v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−ξ2∇⋅(v∇w)+μ2v(1−v−a2u), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−(u+v)w, (x,t)∈Ω×(0,Tmax),∂u∂ν=∂v∂ν=∂w∂ν=0, (x,t)∈∂Ω×(0,Tmax),u(x,0)=u0(x),v(x,0)=v0(x),w(x,0)=w0(x), x∈Ω, | (1.4) |
where ai,ξi,μi>0,i=1,2. Numerous research results have been obtained for such a model. For instance, when the initial value ‖w0‖L∞(Ω) satisfies suitable explicit conditions, Wang et al. [33] elaborated that the system is globally classically solvable. And, they also explored the long-time stability of the system. The global classical solvability of system (1.4) with nonlinear diffusion was discussed in [34]. When removing logistic terms in system (1.4), Zhang and Tao [35] constructed the existence conditions provided that ‖w0‖L∞(Ω)≤√2nπmax{ξ1,ξ2}. Ren and Liu [36] presented the global-in-time existence of weak solutions to the model involving nonlinear chemotactic sensitivity functions under the condition that ‖w0‖≤ˉw with ˉw depending on the coefficients of system. Later, Ren and Liu [37] introduced a definition of weak solutions and showed that these solutions would be smooth after a certain moment T>0.
The forager-exploiter model can sometimes be considered as a variant of chemotaxis-consumption model,
{∂u∂t=Δu+ξ∇⋅(u∇w)+f1(u,v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−χ∇⋅(v∇u)+f2(u,v), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−(u+v)w−μw+r(x,t), (x,t)∈Ω×(0,Tmax), | (1.5) |
where u and v stand for the foragers density and the exploiters density, respectively, w represents the resource concentration, and r(x,t) stands for resource production rate function. Assuming system (1.5) without logistic terms, Winkler [38] provided an explicit condition with respect to r(x,t) and initial data to ensure the global weak solvability. Letting r(x,t)=r0 with some constant r0≥0, Tao and Winkler [39] explored the existence of global classical solutions to this associated system for all suitably regular initial data in one-dimensional space. For spatial dimension n≥2, if the initial data and r(x,t) satisfy some smallness conditions or χ,ξ are small enough, Wang and Wang [40] established the global solvability in the classical sense for the corresponding system. In addition, if f1(u,v)=η1(u−u2) and f2(u,v)=η2(v−v2) with η1,η2>0, Wu and Shen [41] established the global well-posedness under the assumption that θ>(n−2)+n+2 with n≥1. For the case where f1(u,v)=η1u(1−u−a1v) and f2(u,v)=η2v(1−v−a2u) with η1,η2,a1,a2>0, and the third equation of (1.5) is changed with wt=Δw−(u+v)w(1+u+v)θ, Ou and Wang [42] proved the global classical solvability provided that θ>0.
Motivated by the aforementioned works, in the current work, we are concerned with a predator-prey model involving nonlinear nutrient dissipation mechanisms and generalized logistic terms
{∂u∂t=Δu−χ∇⋅(u∇w)+u(a1−b1um−1+c1v), (x,t)∈Ω×(0,Tmax),∂v∂t=Δv−ξ∇⋅(v∇w)+v(a2−b2vl−1−c2u), (x,t)∈Ω×(0,Tmax),∂w∂t=Δw−(uα+vβ)w, (x,t)∈Ω×(0,Tmax),∂u∂ν=∂v∂ν=∂w∂ν=0, (x,t)∈∂Ω×(0,Tmax),u(x,0)=u0(x),v(x,0)=v0(x),w(x,0)=w0(x), x∈Ω, | (1.6) |
with homogeneous Neumann conditions ∂u∂ν=∂v∂ν=∂w∂ν=0 on ∂Ω, where the boundary Ω⊂Rn(n≥2) is smooth, ν is the outward normal vector on ∂Ω, and the parameters ai,bi,ci,α, β,χ,ξ>0 and m,l>1 with i=1,2. The purpose of the current paper is to provide a sufficient condition on global solvability in the classical sense to system (1.6). For this purpose, suppose that the initial values u0,v0, and w0 fulfill
u0,v0,w0∈W2,∞(Ω) with u0,v0,w0≥0,≢0 in Ω. | (1.7) |
We state the main result as follows.
Theorem 1.1. Let n≥2, ai,bi,ci,α, β,χ,ξ>0 and m,l>1 with i=1,2. Suppose that u0,v0, and w0 satisfy (1.7). If m>max{α(n+2)2,1} and l>max{β(n+2)2,1}, then model (1.6) possesses a nonnegative solution in the sense that
(u,v,w)∈⋂k>n[C0([0,∞);W1,k(Ω))∩C2,1(¯Ω×(0,∞))]3, |
which is uniformly-in-time bounded, namely, we can find C>0 fulfilling
‖u(⋅,t)‖W1,k(Ω)+‖v(⋅,t)‖W1,k(Ω)+‖w(⋅,t)‖W1,k(Ω)≤C |
for all k>n and t>0.
Comparing to the linear system explored in [33,35,36], in our conclusion we removed the dependence on the smallness condition of ‖w0‖L∞(Ω), and showed that the existence conditions depend only on the exponents m,l,α,β and spatial dimensions n. In addition, the logistic source terms and nonlinear resource consumption considered here are more complicated than those in [42], thus the result established in this paper seems to be more generalized.
The remaining structure is carried out as follows. In Section 2, we provide some preliminary results, and introduce several useful conclusions that will be utilized in the subsequent part. In Section 3, the proof of the main conclusion is presented.
In this part, we introduce some previously established results which will be useful later. We begin with a local existence conclusion to system (1.6), and the proof can be established through the fixed point theory.
Lemma 2.1. Suppose that Ω⊂Rn is a bounded smooth domain with n≥2, and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. For any u0,v0, and w0 satisfying (1.7), system (1.6) is locally-in-time solvable in the sense that
(u,v,w)∈⋂k>n[C0([0,Tmax);W1,k(Ω))∩C2,1(¯Ω×(0,Tmax))]3, |
on [0,Tmax] with Tmax∈(0,+∞] for all k>n. Furthermore, if Tmax<∞, then
lim supt↗Tmax(‖u(⋅,t)‖W1,k(Ω)+‖v(⋅,t)‖W1,k(Ω)+‖w(⋅,t)‖W1,k(Ω))=∞ | (2.1) |
Proof. As done in [43,44], let ψ=(u,v,w)∈R3. Then, system (1.6) can be reformulated as the following triangular system:
{∂ψ∂t=∇⋅(A(ψ)∇ψ)+σ(ψ), (x,t)∈Ω×(0,Tmax),∂ψ∂ν=0, (x,t)∈∂Ω×(0,Tmax),ψ(⋅,0)=(u0,v0,w0), x∈Ω, | (2.2) |
where
A(ψ)=(10−χu01−ξv001) and σ(ψ)=(u(a1−b1um−1+c1v)v(a2−b2vl−1−c2u)−(uα+vβ)w). |
Since the matrix A(ψ) is positive definite for the given initial data, this asserts that system (2.2) is generally parabolic. Then, Theorems 14.4 and 14.6 in [45] are applicable, and there exists a Tmax≥0 such that system (2.2) admits a solution ψ∈⋂k>n[C0([0,Tmax);W1,k(Ω))∩C2,1(¯Ω×(0,Tmax))]3. Finally, the extensibility criterion can be ensured by applying Theorems 15.5 in [45].
Lemma 2.2. (cf. [23,46]) Let Ω be a smooth bounded domain in Rn with n≥1 and any ρ∈C2(¯Ω) with ∂ρ∂ν|∂Ω=0. For any τ>0 and k>1, there exists C=C(τ,k,Ω)>0 such that
∫∂Ω|∇ρ|2k−2∂|∇ρ|2∂ν≤τ∫Ω|∇ρ|2k−2|D2ρ|2+C∫Ω|∇ρ|2k, | (2.3) |
and
∫Ω|∇ρ|2k+2≤2(4k2+n)‖ρ‖2L∞(Ω)∫Ω|∇ρ|2k−2|D2ρ|2. | (2.4) |
Lemma 2.3. (cf. [40,47]) For some m1,m2>0 and μ=min{1,˜T2} with ˜T∈(0,∞], let z∈C([0,˜T))∩C1((0,˜T)) and y∈L1loc([0,˜T)) be nonnegative such that
dzdt+m1z≤y, t∈(0,˜T) |
and
∫t+μty(s)ds≤m2, t∈(0,˜T−μ). |
Then, there holds
z(t)≤z(0)+2m2+m2m1, t∈(0,˜T). |
This section is dedicated to proving the main conclusion of the paper.
Lemma 3.1. Let n≥2, and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. Then, there exist K0,K1,K2>0 such that
‖w‖L∞(Ω)≤K0, t∈(0,Tmax) | (3.1) |
and
∫Ω(u+v)≤K1, t∈(0,Tmax), | (3.2) |
as well as
∫t+δt∫Ω(um+vl)≤K2, t∈(0,Tmax−δ), | (3.3) |
where δ=min{1,Tmax2}.
Proof. The parabolic comparison principle enables us to obtain (3.1) from the third equation of system (1.6). Next, combining the first and second equations of (1.6), it is not hard to get
ddt∫Ω(c2u+c1v)=a1c2∫Ωu+a2c1∫Ωv−b1c2∫Ωum−b2c1∫Ωvl, t∈(0,Tmax). | (3.4) |
For m,l>1, invoking Young's inequality, one may derive
−b1c2∫Ωum≤−(a1c2+c2)∫Ωu+C1 | (3.5) |
and
−b2c1∫Ωvl≤−(a2c1+c1)∫Ωv+C2, t∈(0,Tmax), | (3.6) |
with some C1,C2>0. Collecting (3.4)–(3.6), one may deduce
ddt∫Ω(c2u+c1v)+∫Ω(c2u+c1v)≤C1+C2, t∈(0,Tmax). | (3.7) |
Applying the ODE comparison principle to inequality (3.7), one can conclude (3.2) directly. Furthermore, integrating both sides of (3.4) from t to t+δ, we can obtain
∫t+δt∫Ω(c2∂u∂t+c1∂v∂t)=∫t+δt∫Ω(a1c2u+a2c1v)−∫t+δt∫Ω(b1c2um+b2c1vl), | (3.8) |
with δ=min{1,Tmax2}. Based on the proven conclusion in (3.2), one may see that
∫t+δt∫Ω(b1c2um+b2c1vl)≤∫t+δt∫Ω(a1c2u+a2c1v)+∫Ω(c2u+c1v)≤C3 | (3.9) |
for all t∈(0,Tmax−δ). Thus, we finish the proof of this lemma.
Lemma 3.2. Let n≥2, and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. For any k>1, there exist K3,K4,K5>0 satisfying
12kddt∫Ω|∇w|2k+∫Ω|∇w|2k≤K3∫Ωuα(k+1)+K4∫Ωvβ(k+1)+K5, t∈(0,Tmax). | (3.10) |
Proof. Due to ∇w⋅∇Δw=12Δ|∇w|2−|D2w|2, we deal with the third equation in (1.6) to deduce
∇w⋅∇wt=∇w⋅∇Δw−∇w⋅∇(uαw+vβw)=12Δ|∇w|2−|D2w|2−∇w⋅∇(uαw+vβw). | (3.11) |
For any k>1, we can obtain from (3.11) that
12kddt∫Ω|∇w|2k+∫Ω|∇w|2k−2|D2w|2+∫Ω|∇w|2k=12∫Ω|∇w|2k−2Δ|∇w|2+∫Ω|∇w|2k−∫Ω|∇w|2k−2∇w⋅∇(uαw+vβw)=H1+H2, | (3.12) |
where H1=12∫Ω|∇w|2k−2Δ|∇w|2+∫Ω|∇w|2k and H2=−∫Ω|∇w|2k−2∇w⋅∇(uαw+vβw). Due to the boundedness of ‖w‖L∞(Ω) in (3.1), we employ (2.4) in Lemma 2.2 to get
∫Ω|∇w|2k+2≤C1∫Ω|∇w|2k−2|D2w|2, t∈(0,Tmax), | (3.13) |
where C1=2(4k2+n)K20>0. In view of (2.3) in Lemma 2.2 and (3.13), it is not hard to deduce from Young's inequality that
H1=12∫∂Ω|∇w|2k−2∂|∇w|2∂ν−12∫Ω∇|∇w|2k−2⋅∇|∇w|2+∫Ω|∇w|2k≤14∫Ω|∇w|2k−2|D2w|2+C2∫Ω|∇w|2k−k−12∫Ω|∇w|2k−4|∇|∇w|2|2≤14∫Ω|∇w|2k−2|D2w|2+14C1∫Ω|∇w|2k+2+C3≤12∫Ω|∇w|2k−2|D2w|2+C3, t∈(0,Tmax), | (3.14) |
with C2>0 and C3=(4C1)kCk+12|Ω|>0. Applying the inequality |Δw|≤√n|D2w|, it can be inferred from (3.1) and integration by parts that
H2=−∫Ω|∇w|2k−2∇w⋅∇(uαw+vβw)=∫Ω(uαw+vβw)∇⋅(∇w|∇w|2k−2)=∫Ω(uαw+vβw)(Δw|∇w|2k−2+(2k−2)|∇w|2k−2|D2w|)≤C4∫Ω(uα+vβ)|∇w|2k−2|D2w|, t∈(0,Tmax), | (3.15) |
where C4=(√n+2(k−1))K0>0. Using (3.13) once more, we see
C4∫Ω(uα+vβ)|∇w|2k−2|D2w|≤14∫Ω|∇w|2k−2|D2w|2+C5∫Ω(u2α+v2β)|∇w|2k−2≤14∫Ω|∇w|2k−2|D2w|2+14C1∫Ω|∇w|2k+2+C6∫Ωuα(k+1)+C6∫Ωvβ(k+1)≤12∫Ω|∇w|2k−2|D2w|2+C6∫Ωuα(k+1)+C6∫Ωvβ(k+1), t∈(0,Tmax), | (3.16) |
with some C5,C6>0. Collecting (3.14), (3.16), and (3.12), for some C7>0, one may get
12kddt∫Ω|∇w|2k+∫Ω|∇w|2k≤C6∫Ωuα(k+1)+C6∫Ωvβ(k+1)+C7, t∈(0,Tmax). | (3.17) |
Therefore, we can obtain (3.10).
Lemma 3.3. Let n≥2 and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. Suppose that for any k>max{(α+β)(n+2)2,1} there is K6>0 satisfying
∫t+δt∫Ω(uαkα+β+vβkα+β)≤K6, t∈(0,T∗max), | (3.18) |
where δ=min{1,Tmax2} and T∗max=Tmax−δ. Then, we can find K7>0 satisfying
‖∇w(⋅,t)‖L2(kα+β−1)(Ω)≤K7, t∈(0,Tmax). | (3.19) |
Proof. Due to Lemma 3.2, it is not hard to find C1,C2,C3>0 satisfying
ddt∫Ω|∇w|2(kα+β−1)+C1∫Ω|∇w|2(kα+β−1)≤C2∫Ω(uαkα+β+vβkα+β)+C3, t∈(0,Tmax). | (3.20) |
Since k>(α+β)(n+2)2, we see that 2(kα+β−1)>n. From (3.18) and Lemma 2.3, it is not difficult to get from (3.20) that
∫Ω|∇w|2(kα+β−1)≤C4, t∈(0,Tmax), | (3.21) |
with some C4>0. Hence, we can conclude (3.17).
Lemma 3.4. Let n≥2, and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. Then, we can find K8,K9>0 to satisfy
‖u(⋅,t)‖L∞(Ω)≤K8 and ‖v(⋅,t)‖L∞(Ω)≤K9, t∈(0,Tmax). | (3.22) |
Proof. Based on the variation-of-constants formula, one may derive
v(⋅,t)=etΔv0−ξ∫t0e(t−s)Δ∇⋅(v∇w)ds+∫t0e(t−s)Δ(a2v−b2vl−c2uv)ds=etΔv0−ξ∫t0e(t−s)Δ∇⋅(v∇w)ds+∫t0e(t−s)Δ[(a2v−b2vl−c2uv)+−(a2v−b2vl−c2uv)−]ds≤etΔv0−ξ∫t0e(t−s)Δ∇⋅(v∇w)ds+∫t0e(t−s)Δ(a2v−b2vl−c2uv)+ds | (3.23) |
for all t∈(0,Tmax). Therefore, one may deduce
‖v(⋅,t)‖L∞(Ω)≤‖etΔv0‖L∞(Ω)+ξ∫t0‖e(t−s)Δ∇⋅(v∇w)‖L∞(Ω)ds+∫t0‖e(t−s)Δ(a2v−b2vl−c2uv)+‖L∞(Ω)ds≤C1+ξ∫t0‖e(t−s)Δ∇⋅(v∇w)‖L∞(Ω)ds+∫t0‖e(t−s)Δ(a2v−b2vl−c2uv)+‖L∞(Ω)ds | (3.24) |
for all t∈(0,Tmax) with some C1>0. From Lemma 3.3, for any k>max{(α+β)(n+2)2,1}, there holds
‖∇w(⋅,t)‖L2(kα+β−1)(Ω)≤K7, t∈(0,Tmax). | (3.25) |
Define κ>0 satisfying n<κ<2(kα+β−1). Let γ=2(kα+β−1)κ2(kα+β−1)−κ>n. Invoking Hölder's inequality and the Lk-interpolation inequality, we conclude from the regularization properties of the Neumann heat semigroup (etΔ)t≥0 (see [48]) that
ξ∫t0‖e(t−s)Δ∇⋅(v∇w)‖L∞(Ω)ds≤C2∫t0(1+(t−s)−12−n2κ)e−λ(t−s)‖(v∇w)‖Lκ(Ω)ds≤C2∫t0(1+(t−s)−12−n2κ)e−λ(t−s)‖v‖Lγ(Ω)‖∇w‖L2(kα+β−1)(Ω)ds≤C3∫t0(1+(t−s)−12−n2κ)e−λ(t−s)‖v‖1γL1(Ω)‖v‖γ−1γL∞(Ω)ds≤C3K1γ1∫t0(1+(t−s)−12−n2κ)e−λ(t−s)‖v‖γ−1γL∞(Ω)ds | (3.26) |
for all t∈(0,Tmax), with some λ,C2,C3>0. Let
I(t)=sups∈(0,t)‖v(⋅,s)‖L∞(Ω), t∈(0,Tmax). | (3.27) |
Due to n<κ<2(kα+β−1), we infer that 1γ∈(0,1) and
∫t0(1+(t−s)−12−n2κ)e−λ(t−s)ds<∞. | (3.28) |
Thus, it can be deduced from (3.26)–(3.28) that
ξ∫t0‖e(t−s)Δ∇⋅(v∇w)‖L∞(Ω)ds≤C4K1γ1Iγ−1γ(t), t∈(0,Tmax), | (3.29) |
with some C4>0. Letting f(y)=a2y−b2yl, due to u,v≥0 and l>1, we know that
(a2v−b2vl−c2uv)+≤(a2v−b2vl)+≤f((a2lb2)1l−1), | (3.30) |
which implies
∫t0‖e(t−s)Δ(a2v−b2vl−c2uv)+‖L∞(Ω)ds≤C5∫t0e−λ(t−s)‖(a2v−b2vl)+‖L∞(Ω)ds≤C6, t∈(0,Tmax), | (3.31) |
where C5,C6>0. Substituting (3.27), (3.29), and (3.31) into (3.24), it can be concluded from Young's inequality that
I(t)≤C1+C4K1γ1Iγ−1γ(t)+C6≤C7+12I(t), t∈(0,Tmax), | (3.32) |
with some C7>0. Therefore, from the definition of I(t), there holds
‖v(⋅,t)‖L∞(Ω)<K8, t∈(0,Tmax), | (3.33) |
with some K8>0. In addition, based on the variation-of-constants formula, we can also obtain
u(⋅,t)=etΔu0−χ∫t0e(t−s)Δ∇⋅(u∇w)ds+∫t0e(t−s)Δ(a1u−b1um+c1uv)ds |
for all t∈(0,Tmax). Due to the L∞−boundedness of v as in (3.33), we derive from m>1 that
(a1u−b1um+c1uv)+≤(a1u−b1um+c1C8u)+≤K9 | (3.34) |
for all t∈(0,Tmax) with some K9>0. Similarly, we can use the same procedures as above to deduce the L∞−boundedness of u. Thus, we finish the proof.
Lemma 3.5. Let n≥2 and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. Then, for any k>1, we can find K10>0 satisfying
12kddt∫Ω(|∇u|2k+|∇v|2k)+∫Ω(|∇u|2k+|∇v|2k)≤K10∫Ω|∇w|2k+2+K10∫Ω|Δw|k+1+K10. |
Proof. Applying the same steps as in (3.11) and (3.12), we conclude from the second equation of system (1.6) that
12kddt∫Ω|∇v|2k+∫Ω|∇v|2k−2|D2v|2+∫Ω|∇v|2k=12∫Ω|∇v|2k−2Δ|∇v|2+ξ∫Ω∇⋅(|∇v|2k−2∇v)(∇v⋅∇w+vΔw)−∫Ω|∇v|2k−2∇v⋅∇(b2vl+c2uv)+(a2+1)∫Ω|∇v|2k=I1+I2+I3+(a2+1)∫Ω|∇v|2k, t∈(0,Tmax), | (3.35) |
where the identity ∇v⋅∇Δv=12Δ|∇v|2−|D2v|2 has been used. Using similar steps as in deriving H1 in Lemma 3.2, we can find C1>0 such that
I1=12∫Ω|∇v|2k−2Δ|∇v|2≤18∫Ω|∇v|2k−2|D2v|2+C1, t∈(0,Tmax). | (3.36) |
For the term I2, we can calculate that
I2=ξ∫Ω∇⋅(|∇v|2k−2∇v)(∇v⋅∇w+vΔw)=ξ∫Ω(∇|∇v|2k−2⋅∇v)(∇v⋅∇w)+ξ∫ΩvΔw(∇|∇v|2k−2⋅∇v)+ξ∫Ω|∇v|2k−2Δv(∇v⋅∇w)+ξ∫Ωv|∇v|2k−2ΔvΔw, t∈(0,Tmax). | (3.37) |
From Lemma 2.2 and (3.18), for some C2>0 we have
∫Ω|∇v|2k+2≤C2∫Ω|∇v|2k−2|D2v|2, t∈(0,Tmax). | (3.38) |
In the following, we shall estimate each term of (3.37). For the first term, we infer from Young's inequality and (3.38) that
ξ∫Ω(∇|∇v|2k−2⋅∇v)(∇v⋅∇w)=ξ(k−1)∫Ω|∇v|2k−4(∇|∇v|2⋅∇v)(∇v⋅∇w)≤2ξ(k−1)∫Ω|∇v|2k−1|D2v||∇w|≤116∫Ω|∇v|2k−2|D2v|2+16ξ2(k−1)2∫Ω|∇v|2k|∇w|2≤116∫Ω|∇v|2k−2|D2v|2+116C2∫Ω|∇v|2k+2+C3∫Ω|∇w|2k+2≤18∫Ω|∇v|2k−2|D2v|2+C3∫Ω|∇w|2k+2, | (3.39) |
with some C3>0. For the second term, we see
ξ∫ΩvΔw(∇|∇v|2k−2⋅∇v)=ξ(k−1)∫Ωv|∇v|2k−4Δw(∇|∇v|2⋅∇v)=2ξ(k−1)∫Ωv|∇v|2k−4Δw((D2v⋅∇v)⋅∇v)≤C4∫Ω|∇v|2k−2|D2v||Δw|, t∈(0,Tmax), | (3.40) |
with some C4>0. Based on Young's inequality and (3.38), the third term can be estimated as
ξ∫Ω|∇v|2k−2Δv(∇v⋅∇w)≤√nξ∫Ω|∇v|2k−1|D2v||∇w|≤116∫Ω|∇v|2k−2|D2v|2+C5∫Ω|∇v|2k|∇w|2≤116∫Ω|∇v|2k−2|D2v|2+116C2∫Ω|∇v|2k+2+C6∫Ω|∇w|2k+2≤18∫Ω|∇v|2k−2|D2v|2+C6∫Ω|∇w|2k+2, t∈(0,Tmax), | (3.41) |
with some C5,C6>0. For the last term, due to (3.22), we have
ξ∫Ωv|∇v|2k−2ΔvΔw≤√nξ∫Ωv|∇v|2k−2|D2v||Δw|≤C7∫Ω|∇v|2k−2|D2v||Δw| | (3.42) |
for all t∈(0,Tmax), with C7>0. From the nonnegativity of u and v, we can obtain
I3=−∫Ω|∇v|2k−2∇v⋅∇(b2vl+c2uv)=−b2l∫Ωvl−1|∇v|2k−c2∫Ωu|∇v|2k−c2∫Ωv|∇v|2k−2∇v⋅∇u≤c2∫Ω|∇v|2k−1|∇u|≤C8∫Ω|∇v|2k+C9∫Ω|∇u|2k, t∈(0,Tmax), | (3.43) |
with some C8,C9>0. By employing Young's inequality, for some C10,C11>0, one may get
C8∫Ω|∇v|2k≤18C2∫Ω|∇v|2k+2+C10≤18∫Ω|∇v|2k−2|D2v|2+C10 | (3.44) |
and
C9∫Ω|∇u|2k≤18C2∫Ω|∇u|2k+2+C11≤18∫Ω|∇u|2k−2|D2u|2+C11 | (3.45) |
for all t∈(0,Tmax). By adding up (3.40) and (3.42), for some C12,C13>0, we can further obtain
ξ∫ΩvΔw(∇|∇v|2k−2⋅∇v)+ξ∫Ωv|∇v|2k−2ΔvΔw≤(C4+C7)∫Ω|∇v|2k−2|D2v||Δw|≤18∫Ω|∇v|2k−2|D2v|2+C12∫Ω|∇v|2k−2|Δw|2≤18∫Ω|∇v|2k−2|D2v|2+14C2∫Ω|∇v|2k+2+C13∫Ω|Δw|k+1≤38∫Ω|∇v|2k−2|D2v|2+C13∫Ω|Δw|k+1, t∈(0,Tmax). | (3.46) |
Thus, we can obtain from (3.35), (3.36), (3.39), (3.41), and (3.44)–(3.46) that
12kddt∫Ω|∇v|2k+18∫Ω|∇v|2k−2|D2v|2+∫Ω|∇v|2k≤C14∫Ω|∇w|2k+2+C13∫Ω|Δw|k+1+18∫Ω|∇u|2k−2|D2u|2+C15, t∈(0,Tmax), | (3.47) |
with some C14,C15>0. Additionally, employing the same derivation processes as above, we can also obtain from the first equation in (1.6) that
12kddt∫Ω|∇u|2k+18∫Ω|∇u|2k−2|D2u|2+∫Ω|∇u|2k≤C16∫Ω|∇w|2k+2+C17∫Ω|Δw|k+1+18∫Ω|∇v|2k−2|D2v|2+C18, t∈(0,Tmax), | (3.48) |
with some C16,C17,C18>0. Thus, the desired conclusion can be deduced by adding up (3.47) and (3.48).
Lemma 3.6. Let n≥2 and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2 and k>max{(α+β)(n+2)2,1}, δ=min{1,Tmax2}, and T∗max=Tmax−δ. Then, we can obtain
‖∇u(⋅,t)‖L2(kα+β−1)(Ω)+‖∇v(⋅,t)‖L2(kα+β−1)(Ω)≤K11, t∈(0,Tmax), | (3.49) |
with some K11>0.
Proof. Set
h(x,t)=−(uα+vβ)w, (x,t)∈Ω×(0,Tmax). | (3.50) |
From the boundedness of ‖w‖L∞(Ω) and (3.18), for δ=min{1,Tmax2} and T∗max=Tmax−δ, we infer that
∫t+δt∫Ω|h|kα+β≤K0∫t+δt∫Ω(uα+vβ)kα+β≤C1∫t+δt∫Ω(uαkα+β+vβkα+β)+C2≤C3 | (3.51) |
for all t∈(0,T∗max), with Ci>0,i=1,...,3. Let w solve the problem
{∂w∂t=Δw+h(x,t), (x,t)∈Ω×(0,Tmax),∂w∂ν=0, (x,t)∈Ω×(0,Tmax),w(x,0)=w0, x∈Ω. | (3.52) |
Thus, we deduce from (3.51) and [49, Lemma 2.5] that
∫t+δt∫Ω|Δw|kα+β≤C4, t∈(0,T∗max), | (3.53) |
with some C4>0. Replacing k in Lemma 3.5 with kα+β−1, we have
12(kα+β−1)ddt∫Ω(|∇u|2(kα+β−1)+|∇v|2(kα+β−1))+∫Ω(|∇u|2(kα+β−1)+|∇v|2(kα+β−1))≤K10∫Ω|∇w|2kα+β+K10∫Ω|Δw|kα+β+K10, t∈(0,Tmax). | (3.54) |
Invoking the Gagliardo-Nirenberg inequality (see [50,51]) and Lemma 3.1, for some C5,C6>0, it is not difficult to get
∫Ω|∇w|2kα+β=‖∇w‖2kα+βL2kα+β(Ω)≤C5‖Δw‖kα+βLkα+β(Ω)‖w‖kα+βL∞(Ω)+C5‖w‖2kα+βL∞(Ω)≤C6∫Ω|Δw|kα+β+C6, t∈(0,Tmax). | (3.55) |
Substituting (3.55) into (3.54), we get
12(kα+β−1)ddt∫Ω(|∇u|2(kα+β−1)+|∇v|2(kα+β−1))+∫Ω(|∇u|2(kα+β−1)+|∇v|2(kα+β−1))≤C7∫Ω|Δw|kα+β+C8, | (3.56) |
with some C7,C8>0. Using Lemma 2.3, we deduce from (3.53) and (3.56) that
∫Ω(|∇u|2(kα+β−1)+|∇v|2(kα+β−1))≤C9, t∈(0,Tmax), | (3.57) |
with some C9>0. Thus, we can deduce (3.49).
Lemma 3.7. Suppose that for any k>max{(α+β)(n+2)2,1}, there is C>0 satisfying
∫t+δt∫Ω(uαkα+β+vβkα+β)≤C, t∈(0,T∗max), | (3.58) |
with δ=min{1,Tmax2} and T∗max=Tmax−δ, then Tmax=∞.
Proof. Due to Lemmas 3.3 and 3.6, it is not difficult to find ˉk=2(kα+β−1)>n and C1>0 satisfying
‖u(⋅,t)‖W1,ˉk(Ω)+‖v(⋅,t)‖W1,ˉk(Ω)+‖w(⋅,t)‖W1,ˉk(Ω)≤C1, t∈(0,T∗max). | (3.59) |
Thus, based on Lemma 2.1, we know Tmax=∞.
The proof of Theorem 1.1 Let n≥2 and ai,bi,ci,α,β,χ,ξ>0,m,l>1 with i=1,2. We see that if m>max{α(n+2)2,1} and l>max{β(n+2)2,1}, Theorem 1.1 can be concluded from Lemma 3.7 and (3.3).
In this paper, we consider a predator-prey model involving nonlinear nutrient dissipation mechanisms and generalized logistic terms, and the sufficient condition for system (1.6) to have global solvability in the classical sense has been found. Compared to previous work, we use a method of a series of bootstrap-type arguments for some variational structures to obtain the global classical solvability of the system, overcoming the problems caused by nonlinear terms. The novelty of this paper lies in the fact that the existence result established here is more generalized depending only on the nonlinear power exponents and spatial dimensions.
From a purely mathematical perspective, there are also other interesting questions related to system (1.6) that are worth further exploration. For example, by adjusting parameters such as ai,bi, and ci, it can exhibit richer dynamic behaviors, such as oscillation, stable equilibrium, and bifurcation, so as to adapt to different practical problems. We will consider these issues in our future work.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We would like to thank the anonymous referees for many useful comments and suggestions that greatly improve the work. This work was partially supported by the Natural Science Foundation of Henan Province No. 242300421695 and Nanhu Scholars Program for Young Scholars of XYNU No. 2020017.
The authors declare that there is no conflict of interest.
[1] |
E. Keller, L. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol., 26 (1970), 399–415. https://doi.org/10.1016/0022-5193(70)90092-5 doi: 10.1016/0022-5193(70)90092-5
![]() |
[2] |
C. Wang, P. Wang, X. Zhu, Global dynamics in a chemotaxis system involving nonlinear indirect signal secretion and logistic source, Z. Angew. Math. Phys., 74 (2023), 237. https://doi.org/10.1007/s00033-023-02126-2 doi: 10.1007/s00033-023-02126-2
![]() |
[3] |
M. Winkler, Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source, Commun. Partial Differ. Equations, 35 (2010), 1516–1537. https://doi.org/10.1080/03605300903473426 doi: 10.1080/03605300903473426
![]() |
[4] |
C. Lo, J. Rodrigues, Global existence for nonlocal quasilinear diffusion systems in nonisotropic nondivergence form, Math. Nachr., 297 (2024), 2122–2147. https://doi.org/10.1002/mana.202200250 doi: 10.1002/mana.202200250
![]() |
[5] |
M. Winkler, Does a "volume-filling effect" always prevent chemotactic collapse?, Math. Methods Appl. Sci., 33 (2010), 12–24. https://doi.org/10.1002/mma.1146 doi: 10.1002/mma.1146
![]() |
[6] |
C. Wang, J. Zhu, Blow-up analysis to a quasilinear chemotaxis system with nonlocal logistic effect, Bull. Malays. Math. Sci. Soc., 47 (2024), 60. https://doi.org/10.1007/s40840-024-01659-7 doi: 10.1007/s40840-024-01659-7
![]() |
[7] |
M. Winkler, How far do chemotaxis-driven forces influence regularity in the Navier-Stokes system?, Trans. Am. Math. Soc., 369 (2017), 3067–3125. https://doi.org/10.1090/tran/6733 doi: 10.1090/tran/6733
![]() |
[8] |
J. Zheng, An optimal result for global existence and boundedness in a three-dimensional Keller-Segel-Stokes, J. Differ. Equations, 267 (2019), 2385–2415. https://doi.org/10.1016/j.jde.2019.03.013 doi: 10.1016/j.jde.2019.03.013
![]() |
[9] |
W. Zhang, Global solutions in a chemotaxis consumption model with singular sensitivity, Z. Angew. Math. Phys., 74 (2023), 165. https://doi.org/10.1007/s00033-023-02049-y doi: 10.1007/s00033-023-02049-y
![]() |
[10] |
H. Liu, C. Lo, Determining a parabolic system by boundary observation of its non-negative solutions with biological applications, Inverse Probl., 40 (2024), 025009. https://doi.org/10.1088/1361-6420/ad149f doi: 10.1088/1361-6420/ad149f
![]() |
[11] |
N. Bellomo, A. Bellouquid, Y. Tao, M. Winkler, Toward a mathematical theory of Keller-Segel models of pattern formation in biological tissues, Math. Models Methods Appl. Sci., 25 (2015), 1663–1763. https://doi.org/10.1142/S021820251550044X doi: 10.1142/S021820251550044X
![]() |
[12] |
P. Liu, J. Shi, Z. 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
![]() |
[13] |
G. Fragnelli, Positive periodic solutions for a system of anisotropic parabolic equations, J. Math. Anal. Appl., 367 (2010), 204–228. https://doi.org/10.1016/j.jmaa.2009.12.039 doi: 10.1016/j.jmaa.2009.12.039
![]() |
[14] |
H. Alaa, N. Alaa, A. Bouchriti, A. Charkaoui, An improved nonlinear anisotropic model with p(x)-growth conditions applied to image restoration and enhancement, Math. Meth. Appl. Sci., 47 (2024), 7546–7575. https://doi.org/10.1002/mma.9989 doi: 10.1002/mma.9989
![]() |
[15] |
A. Charkaoui, A. Ben-Loghfyry, S. Zeng, A novel parabolic model driven by double phase flux operator with variable exponents: Application to image decomposition and denoising, Comput. Math. Appl., 174 (2024), 97–141. https://doi.org/10.1016/j.camwa.2024.08.021 doi: 10.1016/j.camwa.2024.08.021
![]() |
[16] |
A. Charkaoui, A. Ben-Loghfyry, S. Zeng, Nonlinear parabolic double phase variable exponent systems with applications in image noise removal, Appl. Math. Modell., 132 (2024), 495–530. https://doi.org/10.1016/j.apm.2024.04.059 doi: 10.1016/j.apm.2024.04.059
![]() |
[17] |
J. Ahn, M. Winkler, A critical exponent for blow-up in a two-dimensional chemotaxis-consumption system, Calc. Var. Partial Differ. Equations, 62 (2023), 6. https://doi.org/10.1007/s00526-023-02523-5 doi: 10.1007/s00526-023-02523-5
![]() |
[18] |
W. Lyu, Asymptotic stabilization for a class of chemotaxis-consumption systems with generalized logistic source, Nonlinear Anal., 217 (2022), 112737. https://doi.org/10.1016/j.na.2021.112737 doi: 10.1016/j.na.2021.112737
![]() |
[19] |
Y. Wang, M. Winkler, Finite-time blow-up in a repulsive chemotaxis-consumption system, Proc. Roy. Soc. Edinburgh Sect. A, 153 (2023), 1150–1166. https://doi.org/10.1017/prm.2022.39 doi: 10.1017/prm.2022.39
![]() |
[20] |
Y. Tao, Boundedness in a chemotaxis model with oxygen consumption by bacteria, J. Math. Anal. Appl., 381 (2011), 521–529. https://doi.org/10.1016/j.jmaa.2011.02.041 doi: 10.1016/j.jmaa.2011.02.041
![]() |
[21] |
K. Baghaei, A. Khelghati, Boundedness of classical solutions for a chemotaxis model with consumption of chemoattractant, C. R. Math. Acad. Sci. Paris, 355 (2017), 633–639. https://doi.org/10.1002/mma.4264 doi: 10.1002/mma.4264
![]() |
[22] |
M. Fuest, Analysis of a chemotaxis model with indirect signal absorption, J. Differ. Equations, 267 (2019), 4778–4806. https://doi.org/10.1016/j.jde.2019.05.015 doi: 10.1016/j.jde.2019.05.015
![]() |
[23] |
J. Lankeit, Y. Wang, Global existence, boundedness and stabilization in a high-dimensional chemotaxis system with consumption, Discrete Contin. Dyn. Syst., 37 (2017), 6099–6121. https://doi.org/10.3934/dcds.2017262 doi: 10.3934/dcds.2017262
![]() |
[24] |
C. Wang, Z. Zheng, X. Zhu, Dynamic behavior analysis to a generalized chemotaxis-consumption system, J. Math. Phys., 65 (2024), 011503. https://doi.org/10.1063/5.0176530 doi: 10.1063/5.0176530
![]() |
[25] |
Y. Chiyo, S. Frassu, G. Viglialoro, A nonlinear attraction-repulsion Keller-Segel model with double sublinear absorptions: Criteria toward boundedness, Commun. Pure Appl. Anal., 22 (2023), 1783–1809. https://doi.org/10.3934/cpaa.2023047 doi: 10.3934/cpaa.2023047
![]() |
[26] |
C. Wang, Z. Zheng, Global boundedness for a chemotaxis system involving nonlinear indirect consumption mechanism, Discrete Contin. Dyn. Syst. B, 29 (2024), 2141–2157. https://doi.org/10.3934/dcdsb.2023171 doi: 10.3934/dcdsb.2023171
![]() |
[27] |
X. Bai, M. Winkler, Equilibration in a fully parabolic two-species chemotaxis system with competitive kinetics, Indiana Univ. Math. J., 65 (2016), 553–583. https://doi.org/10.1512/iumj.2016.65.5776 doi: 10.1512/iumj.2016.65.5776
![]() |
[28] |
M. Mizukami, Boundedness and asymptotic stability in a two-species chemotaxis-competition model with signal-dependent sensitivity, Discrete Contin. Dyn. Syst. Ser. B, 22 (2017), 2301–2319. https://doi.org/10.3934/dcdsb.2017097 doi: 10.3934/dcdsb.2017097
![]() |
[29] |
M. Mizukami, Improvement of conditions for asymptotic stability in a two-species chemotaxiscompetition model with signal-dependent sensitivity, Discrete Contin. Dyn. Syst. Ser. B, 13 (2020), 269–278. https://doi.org/10.3934/dcdss.2020015 doi: 10.3934/dcdss.2020015
![]() |
[30] |
K. Lin, C. Mu, Convergence of global and bounded solutions of a two-species chemotaxis model with a logistic source, Discrete Contin. Dyn. Syst. Ser. B, 22 (2017), 2233–2260. https://doi.org/10.3934/dcdsb.2017094 doi: 10.3934/dcdsb.2017094
![]() |
[31] |
K. Lin, C. Mu, L. Wang, Boundedness in a two-species chemotaxis system, Math. Methods Appl. Sci., 38 (2015), 5085–5096. https://doi.org/10.1002/mma.3429 doi: 10.1002/mma.3429
![]() |
[32] |
L. Miao, H. Yang, S. Fu, Global boundedness in a two-species predator-prey chemotaxis model, Appl. Math. Lett., 111 (2021), 106–639. https://doi.org/10.1016/j.aml.2020.106639 doi: 10.1016/j.aml.2020.106639
![]() |
[33] |
L. Wang, C. Mu, X. Hu, P. Zheng, Boundedness and asymptotic stability of solutions to a two-species chemotaxis system with consumption of chemoattractant, J. Differ. Equations, 264 (2018), 3369–3401. https://doi.org/10.1016/j.jde.2017.11.019 doi: 10.1016/j.jde.2017.11.019
![]() |
[34] |
J. Zhang, X. Hu, L. Wang, L. Qu, Boundedness in a quasilinear two-species chemotaxis system with consumption of chemoattractant, Electron. J. Qual. Theory Differ. Equations, 31 (2019), 1–12. https://doi.org/10.14232/ejqtde.2019.1.31 doi: 10.14232/ejqtde.2019.1.31
![]() |
[35] |
Q. Zhang, W. Tao, Boundedness and stabilization in a two-species chemotaxis system with signal absorption, Comput. Math. Appl., 78 (2019), 2672–2681. https://doi.org/10.1016/j.camwa.2019.04.008 doi: 10.1016/j.camwa.2019.04.008
![]() |
[36] |
G. Ren, B. Liu, Global existence and asymptotic behavior in a two-species chemotaxis system with logistic source, J. Differ. Equations, 269 (2020), 1484–1520. https://doi.org/10.1016/j.jde.2020.01.008 doi: 10.1016/j.jde.2020.01.008
![]() |
[37] |
G. Ren, B. Liu, Global solvability and asymptotic behavior in a two-species chemotaxis system with Lotka-Volterra competitive kinetics, Math. Models Methods Appl. Sci., 31 (2021), 941–978. https://doi.org/10.1142/S0218202521500238 doi: 10.1142/S0218202521500238
![]() |
[38] |
M. Winkler, Global generalized solutions to a multi-dimensional doubly tactic resource consumption model accounting for social interactions, Math. Models Methods Appl. Sci., 29 (2019), 373–418. https://doi.org/10.1142/S021820251950012X doi: 10.1142/S021820251950012X
![]() |
[39] |
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
![]() |
[40] |
J. Wang, M. Wang, Global bounded solution of the higher-dimensional forager-exploiter model with/without growth sources, Math. Models Methods Appl. Sci., 30 (2020), 1297–1323. https://doi.org/10.1142/S0218202520500232 doi: 10.1142/S0218202520500232
![]() |
[41] |
D. Wu, S. Shen, Global boundedness and stabilization in a forager-exploiter model with logistic growth and nonlinear resource consumption, Nonlinear Anal. Real World Appl., 72 (2023), 103–854. https://doi.org/10.1016/j.nonrwa.2023.103854 doi: 10.1016/j.nonrwa.2023.103854
![]() |
[42] |
H. Ou, L. Wang, Boundedness in a two-species chemotaxis system with nonlinear resource consumption, Qual. Theory Dyn. Syst., 23 (2024), 14. https://doi.org/10.1007/s12346-023-00873-1 doi: 10.1007/s12346-023-00873-1
![]() |
[43] |
Y. Choi, Z. Wang, Prevention of blow-up by fast diffusion in chemotaxis, J. Math. Anal. Appl., 362 (2010), 553–564. https://doi.org/10.1016/j.jmaa.2009.08.012 doi: 10.1016/j.jmaa.2009.08.012
![]() |
[44] |
M. Delgado, I. Gayte, C. Morales-Rodrigo, A. Suárez, An angiogenesis model with nonlinear chemotactic response and flux at the tumor boundary, Nonlinear Anal., 72 (2010), 330–347. https://doi.org/10.1016/j.na.2009.06.057 doi: 10.1016/j.na.2009.06.057
![]() |
[45] | H. Schmeisser, H. Triebel, Function spaces, differential operators and nonlinear analysis, in Teubner-Texte zur Mathematik, 133 (1993), 9–126. https://doi.org/10.1007/978-3-663-11336-2 |
[46] |
J. Wang, Global existence and boundedness of a forager-exploiter system with nonlinear diffusions, J. Differ. Equations, 276 (2021), 460–492. https://doi.org/10.1016/j.jde.2020.12.028 doi: 10.1016/j.jde.2020.12.028
![]() |
[47] |
C. Stinner, C. Surulescu, M. Winkler, Global weak solutions in a PDE-ODE system modeling multiscale cancer cell invasion, SIAM J. Math. Anal., 46 (2014), 1969–2007. https://doi.org/10.1137/13094058X doi: 10.1137/13094058X
![]() |
[48] |
M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segal 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
![]() |
[49] |
J. Wang, Global existence and stabilization in a forager-exploiter model with general logistic sources, Nonlinear Anal., 222 (2022), 112–985. https://doi.org/10.1016/j.na.2022.112985 doi: 10.1016/j.na.2022.112985
![]() |
[50] | A. Friedman, Partial Different Equations, Holt Rinehart and Winston, New York, 1969. |
[51] | L. Nirenberg, On elliptic partial differential equations, Ann. Sc. Norm. Super. Pisa Sci. Fis. Mat., III. Ser., 13 (1959), 115–162. |