This paper deals with a two-species chemotaxis-competition system involving singular sensitivity and indirect signal production:
{ut=∇⋅(D(u)∇u)−χ1∇⋅(uzk∇z)+μ1u(1−u−a1v),x∈Ω, t>0,vt=∇⋅(D(v)∇v)−χ2∇⋅(vzk∇z)+μ2v(1−v−a2u),x∈Ω, t>0,wt=Δw−w+u+v,x∈Ω, t>0,zt=Δz−z+w,x∈Ω, t>0,
where Ω⊂Rn is a convex smooth bounded domain with homogeneous Neumann boundary conditions. The diffusion functions D(u),D(v) are assumed to fulfill D(u)≥(u+1)θ1 and D(v)≥(v+1)θ2 with θ1,θ2>0, respectively. The parameters are k∈(0,12)∪(12,1], χi>0,(i=1,2). Additionally, μi should be large enough positive constants, and ai should be positive constants which are less than the quantities associated with |Ω|. Through constructing some appropriate Lyapunov functionals, we can find the lower bounds of ∫Ωu and ∫Ωv. This suggests that any occurrence of extinction, if it happens, will be localized spatially rather than affecting the population as a whole. Moreover, we demonstrate that the solution remains globally bounded if min{θ1,θ2}>1−2n+1 for n≥2.
Citation: Dongxiu Wang, Fugeng Zeng, Lei Huang, Luxu Zhou. Persistence and boundedness in a two-species chemotaxis-competition system with singular sensitivity and indirect signal production[J]. Mathematical Biosciences and Engineering, 2023, 20(12): 21382-21406. doi: 10.3934/mbe.2023946
[1] | Xu Song, Jingyu Li . Asymptotic stability of spiky steady states for a singular chemotaxis model with signal-suppressed motility. Mathematical Biosciences and Engineering, 2022, 19(12): 13988-14028. doi: 10.3934/mbe.2022652 |
[2] | Jean-Jacques Kengwoung-Keumo . Competition between a nonallelopathic phytoplankton and an allelopathic phytoplankton species under predation. Mathematical Biosciences and Engineering, 2016, 13(4): 787-812. doi: 10.3934/mbe.2016018 |
[3] | Tianqi Song, Chuncheng Wang, Boping Tian . Mathematical models for within-host competition of malaria parasites. Mathematical Biosciences and Engineering, 2019, 16(6): 6623-6653. doi: 10.3934/mbe.2019330 |
[4] | Tianyuan Xu, Shanming Ji, Chunhua Jin, Ming Mei, Jingxue Yin . EARLY AND LATE STAGE PROFILES FOR A CHEMOTAXIS MODEL WITH DENSITY-DEPENDENT JUMP PROBABILITY. Mathematical Biosciences and Engineering, 2018, 15(6): 1345-1385. doi: 10.3934/mbe.2018062 |
[5] | Fu-Yuan Tsai, Feng-BinWang . Mathematical analysis of a chemostat system modeling the competition for light and inorganic carbon with internal storage. Mathematical Biosciences and Engineering, 2019, 16(1): 205-221. doi: 10.3934/mbe.2019011 |
[6] | Wenjie Zhang, Lu Xu, Qiao Xin . Global boundedness of a higher-dimensional chemotaxis system on alopecia areata. Mathematical Biosciences and Engineering, 2023, 20(5): 7922-7942. doi: 10.3934/mbe.2023343 |
[7] | Xiaoli Hu, Shengmao Fu . Global boundedness and stability for a chemotaxis model of Boló’s concentric sclerosis. Mathematical Biosciences and Engineering, 2020, 17(5): 5134-5146. doi: 10.3934/mbe.2020277 |
[8] | Danfeng Pang, Yanni Xiao . The SIS model with diffusion of virus in the environment. Mathematical Biosciences and Engineering, 2019, 16(4): 2852-2874. doi: 10.3934/mbe.2019141 |
[9] | Zhilan Feng, Robert Swihart, Yingfei Yi, Huaiping Zhu . Coexistence in a metapopulation model with explicit local dynamics. Mathematical Biosciences and Engineering, 2004, 1(1): 131-145. doi: 10.3934/mbe.2004.1.131 |
[10] | Chichia Chiu, Jui-Ling Yu . An optimal adaptive time-stepping scheme for solving reaction-diffusion-chemotaxis systems. Mathematical Biosciences and Engineering, 2007, 4(2): 187-203. doi: 10.3934/mbe.2007.4.187 |
This paper deals with a two-species chemotaxis-competition system involving singular sensitivity and indirect signal production:
{ut=∇⋅(D(u)∇u)−χ1∇⋅(uzk∇z)+μ1u(1−u−a1v),x∈Ω, t>0,vt=∇⋅(D(v)∇v)−χ2∇⋅(vzk∇z)+μ2v(1−v−a2u),x∈Ω, t>0,wt=Δw−w+u+v,x∈Ω, t>0,zt=Δz−z+w,x∈Ω, t>0,
where Ω⊂Rn is a convex smooth bounded domain with homogeneous Neumann boundary conditions. The diffusion functions D(u),D(v) are assumed to fulfill D(u)≥(u+1)θ1 and D(v)≥(v+1)θ2 with θ1,θ2>0, respectively. The parameters are k∈(0,12)∪(12,1], χi>0,(i=1,2). Additionally, μi should be large enough positive constants, and ai should be positive constants which are less than the quantities associated with |Ω|. Through constructing some appropriate Lyapunov functionals, we can find the lower bounds of ∫Ωu and ∫Ωv. This suggests that any occurrence of extinction, if it happens, will be localized spatially rather than affecting the population as a whole. Moreover, we demonstrate that the solution remains globally bounded if min{θ1,θ2}>1−2n+1 for n≥2.
Chemotaxis refers to the process by which cells move directionally along a concentration gradients of chemical stimuli [1]. The classic Keller-Segel model, along with its numerous variations, has undergone extensive investigations and analyses by numerous researchers following the groundbreaking work of Keller and Segel. The mechanisms underlying this model, including aspects such as cell diffusion, chemotaxis sensitivity, and cell growth and death, have been deeply explored.
The pioneering system for single-species, single-stimulus chemotaxis was as follows:
{ut=d1Δu−∇⋅(uχ(v)∇v)+f(u),x∈Ω, t>0,vt=d2Δv−v+u,x∈Ω, t>0,∂u∂ν=∂v∂ν=0,x∈∂Ω, t>0,(u,v)(x,0)=(u0,v0)(x),x∈Ω, | (1.1) |
where f(u) represents logistic sources. If χ(v)=χ is a positive constant, then d1=1,d2=1, and f(u)=0; this is the most primitive chemotactic model presented by Keller and Segel [2], dating back to 1970. In this case, the solution of the system (1.1), especially in high dimensional space, may exhibit a blow-up phenomenon in either finite or infinite time [3]. However, in the event of the logistic source f(u)=μ0u−μ1u2, Winkler has demonstrated that it possesses globally bounded solutions in high-dimensional systems for a sufficiently large μ1>0 [4]. Additionally, it has a classical solution in three dimensions for any μ1>0, provided that μ0 is not too large [5]. The findings indicate that the blow-up phenomenon can be effectively mitigated by implementing a suitable logistic source term. Importantly, these results hold true even when considering the conditions where d1>0,d2>0 [6]. Additionally, Tao and Winkler established the mass persistence of system (1.1) by constructing an energy function, which explains the persistence of the population as a whole, and the fact that any extinction must occur within a localized spatial region [7]. Apart from that, if χi(w) is a nonlinear chemotaxis sensitivity function given by χi(v)=χv, then it indicates that the sensitivity to chemotaxis is inversely proportional to the density of the signal function. Furthermore, the global existence and boundedness of classical solutions have been established in [8], as well as the global existence of weak solutions under various conditions. For more comprehensive information regarding the qualitative dynamics of system (1.1) and its variants, please consult [9,10,11] and the references therein.
In the aforementioned systems, the signal are directly generated by the cells themselves; however, in realistic situations, signal production may be indirect or multiple by different mechanisms. For instance, Strohm et al. [12] examined the reproductive and accumulation patterns of mountain pine beetles in forest habitats, where flying mountain pine beetles were attracted towards signals secreted by nesting mountain pine beetles, which served as indirect signals. A single-stimulus chemotaxis system with indirect signal production is presented by the following:
{ut=Δu−∇⋅(uχ(v)∇v)+f(u),x∈Ω, t>0,vt=Δv+h(v,w)x∈Ω, t>0,wt=εΔw−δw+u,x∈Ω, t>0,∂u∂ν=∂v∂ν=∂w∂ν=0,x∈∂Ω, t>0,(u,v,w)(x,0)=(u0,v0,w0)(x),x∈Ω. | (1.2) |
Under the conditions f(u)=ru−μu2, χ(v)=χ and ε=0, Hu et al. [13] investigated the boundedness and exponential convergence of solutions. For the case where χ(v)=χv, ε=1 and δ=1, Xing et al. [14] proved that the solution of the system (1.2) is globally bounded in two dimensions and converges exponentially to the steady state if h(v,w)=−v+w and f(u)=0. For the case where χ(v)=χvk, f(u)=ru−μu2 and ε=δ=1, [15] obtained the global boundedness of classical solution of the system (1.2).
On the other hand, there is often an interaction between multiple populations and multiple chemicals that simultaneously occur in a particular environment, thus resulting in competition among them. Therefore, we proceed to directly introduce the following chemotaxis system that incorporates two-species single-stimulus competitive kinetics without indirect signal production:
{ut=Δu−∇⋅(uχ1(w)∇w)+μ1u(1−u−a1v),x∈Ω, t>0,vt=Δv−∇⋅(vχ2(w)∇w)+μ2v(1−a2u−v),x∈Ω, t>0,τwt=Δw−λw+b1u+b2v,x∈Ω, t>0,∂u∂ν=∂v∂ν=∂w∂ν=0,x∈∂Ω, t>0,(u,v,w)(x,0)=(u0,v0,w0)(x),x∈Ω, | (1.3) |
where λ,ai,bi,μi>0 for i=1,2, τ=0 or 1. For the case where χi(w)=χi>0, if τ=0, the coexisting equilibrium state for n≥1 can be found in [16]. Tell and Winkler [17] proved the global asymptotic stability of system (1.3) in high-dimensional scenarios, thus indicating that species groups can coexist under appropriate conditions. Moreover, some blow-up phenomena of system (1.3) can be found in [18,19]. If τ=1, then the large-time behavior and global existence of the system have been extensively investigated in numerous studies [20,21]. For the case where χi(w)=χiw, Mizukami obtained the asymptotic stability of system (1.3) if a1,a2∈(0,1) under n≥2. Qiu et al. demonstrated in [22] that under appropriate parameter conditions, after substituting wt=Δw−(αu+βv)w into the third equation, the system possesses a unique globally uniform bounded solution. Furthermore, related variations of system (1.3) have been investigated to understand the asymptotic behavior of diffused Lotka-Volterra competition models. For more details, please refer to [23].
Now we are in the position to introduce the two-species chemotaxis-competition system involving indirect signal production:
{ut=d1Δu−∇⋅(uχ1(z)∇z)+μ1u(1−u−a1v),x∈Ω, t>0,vt=d2Δv−∇⋅(vχ2(z)∇z)+μ2v(1−v−a2u),x∈Ω, t>0,τwt=Δw−w+u+v,x∈Ω, t>0,τzt=Δz+h(w,z),x∈Ω, t>0,∂u∂ν=∂v∂ν=∂w∂ν=∂z∂ν=0,x∈∂Ω, t>0,(u,v,w,z)(x,0)=(u0,v0,w0,z0)(x),x∈Ω, | (1.4) |
where τ can be either 0 or 1. For the case where χi(z)=χi, when h(w,z)=−wz and τ=1, the boundedness of solution for μ1,μ2>0 and n≤2 is obtained. Furthermore, the asymptotic stabilization of solutions in two dimensions has been shown if a1,a2∈(0,1) along with a1=1>a2>0 [24]. The boundedness and stabilization of system (1.4) for h(w,z)=−z+w and τ∈{0,1} were derived in [25]. Tu et al. [26] studied the global boundedness and regularity of the classical solution of the system, for τ=1, and the third and fourth equations of the system (1.4) are replaced with wt=Δw−λ1w+α11u+α12v, zt=Δz−λ2z+α21u+α22v. For the case where χi(z)=χiz and τ=1, if 0<max{χ1,χ2}<2n, n≥2 or χi>0, n=2, then system (1.4) exhibits a unique global solution. Furthermore, if τ=0, then the boundedness of solution is also established in two-dimensional systems [27]. For the latest research on the well-posedness behavior for two-competing-species two-stimuli chemotaxis models, please refer to the relevant literature [28,29,30] and their references.
From both mathematical and biological perspectives, it is of great interest to investigate whether populations persist and remain limited in size. To the best of our knowledge, there are few findings on the persistence of mass and boundedness in multi-species and multi-chemicals issues, particularly those involving chemotaxis singular sensitivity functions χi(z)=χizk (k>0). In addition, the nonlinear diffusion functions D(u),D(v) are expected to prevent blow-up solutions in chemotaxis systems [31,32]. Consequently, we deal with the following two-species two-chemicals chemotaxis-competition system involving singular sensitivity and indirect signal production:
{ut=∇⋅(D(u)∇u)−χ1∇⋅(uzk∇z)+μ1u(1−u−a1v),x∈Ω, t>0,vt=∇⋅(D(v)∇v)−χ2∇⋅(vzk∇z)+μ2v(1−v−a2u),x∈Ω, t>0,wt=Δw−w+u+v,x∈Ω, t>0,zt=Δz−z+w,x∈Ω, t>0,∂u∂ν=∂v∂ν=∂w∂ν=∂z∂ν=0,x∈∂Ω, t>0,(u,v,w,z)(x,0)=(u0,v0,w0,z0)(x),x∈Ω, | (1.5) |
which is associated with smooth boundary ∂Ω in a bounded convex domain Ω⊂Rn; the nonlinear diffusion functions D(u) and D(v) satisfy the following:
D(u),D(v)∈C2([0,∞)), | (1.6) |
D(u)≥(u+1)θ1, for all u>0 | (1.7) |
and
D(v)≥(v+1)θ2, for all v>0 | (1.8) |
with θ1,θ2>0. The initial data (u0,v0,w0,z0) satisfy the following:
0<u0∈C0(ˉΩ),0<v0∈C0(ˉΩ),0<w0∈W1,∞(Ω),0<z0∈W1,∞(Ω). | (1.9) |
In this scenario, u and v represent the population densities of two competing species, while w and z denote the concentrations of chemical substances. Importantly, both biological species from the two competing populations are attracted to the same chemical signal z. It is worth noting that z is secreted by w, which in turn is secreted by u and v.
In the current paper, we aim to delve deeper into the fundamental questions mentioned above. The main results are presented as follows.
Theorem 1.1. (Persistence) Consider that Ω⊂Rn(n≥2) is a bounded convex domain with a smooth boundary. Suppose that D(u) and D(v) satisfy (1.6), (1.7) and (1.8). Let the parameters χi>0, 0<a1<|Ω|m2,0<a2<|Ω|m1, k∈(0,12)∪(12,1] and μi=μi(χi,k,ai,Ω,u0,v0,w0,z0)(i=1,2) be large enough. If that the initial data (u0,v0,w0,z0) satisfy (1.9) and for any choice of constants Cw,Cz>0,K>0 and S>0 satisfying
∫Ωu0≤m1, ∫Ωv0≤m2, ∫Ωw20≤Cw, ∫Ωz20≤Cz, |
and
∫Ωlnu0≥−K, ∫Ωlnv0≥−K, ∫Ωlnz0≥−S, |
where m1:=max{∫Ωu0,|Ω|}, m2:=max{∫Ωv0,|Ω|}, then for all t∈(0,Tmax), we can find positive constants mu(m1,m2,Cw,Cz,K,S,χ1,μ1,a1),mv(m1,m2,Cw,Cz,K,S,χ2,μ2,a2) such that
∫Ωu≥mu and ∫Ωv≥mv. |
Remark 1.1. In Section 3, we intricately classify our discussion into three separate cases depending on the value range of k, namely k∈(0,12), k∈(12,1) and k=1. It is worth mentioning that handling the second term on the right-hand side of inequality (3.13), specifically when k=12, poses a significant challenge in finding a suitable differential inequality related to ∫Ω|∇z|2z. Consequently, this aspect remains an open problem that requires further investigation.
Theorem 1.2. (Boundedness) Consider a bounded convex domain Ω⊂Rn(n≥2) with a smooth boundary, and assume that the conditions are the same as those within Theorem 1.1. Moreover, it is required that θ>1−2n+1. Then, system (1.5) has a unique classical solution that remains globally bounded.
Remark 1.2. Based on the result in Theorem 1.1, it becomes feasible to establish a lower bound for z. Since infx∈Ωz(x,t)≥δ with δ independent of t, the system (1.5) has a globally bounded solution.
Lemma 2.1. Consider a bounded domain Ω⊂Rn(n≥1) with a smooth boundary, and the parameters ai,μi,χi(i=1,2) are assumed to be positive. Let the initial data (u0,v0,w0,z0) satisfy condition (1.9). For any q>n, there exist Tmax∈(0,∞] and a unique quadruple (u,v,w,z) of nonnegative functions fulfilling the following:
u∈C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax)),v∈C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax)),w∈C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax))∩L∞loc([0,Tmax);W1,q(Ω)),z∈C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax))∩L∞loc([0,Tmax);W1,q(Ω)), | (2.1) |
which classically solve (1.5) in Ω×[0,Tmax).
Additionally, if Tmax<∞, it follows that
‖u(⋅,t)‖L∞(Ω)+‖v(⋅,t)‖L∞(Ω)+‖w(⋅,t)‖W1,q(Ω)+‖z(⋅,t)‖W1,q(Ω)→∞ as t↗∞. | (2.2) |
Proof. Based on the parabolic regularity theory and the standard contraction mapping argument described in [34,35], the local existence of the classical solution to (1.5) can be similarly derived.
To obtain the upper bound for ∫Ωw2, it is necessary to utilize the following auxiliary lemma that ensures the boundedness of solutions to a linearly damped ordinary differential equation with an inhomogeneity.
Lemma 2.2. [36] Let us consider the assumption where f is a nonnegative absolutely continuous function on [0,τ), and g is a nonnegative function belonging to C0[0,τ). They satisfy the following conditions:
f′(t)+αf(t)≤g(t), a.e. t∈(0,τ), |
∫t+1tg(s)ds≤β, t∈[0,τ−1), |
where α>0 and β>0. Under these assumptions, we can then conclude that for 0<t<τ,
f(t)≤max{f(0)+β,βα+2β}. |
Subsequently, we present a number of well-established findings regarding the lower bound of ∫Ωlnψ in accordance with ∫Ω|∇ψ|2ψ2, where ψ∈C1(ˉΩ) is positive. Additionally, the results provide quantitative information regarding the magnitude of the point set and a variation of the Poincaré inequality. For detailed proofs of these results, we refer to [7].
Lemma 2.3. [7] Let α,β>0 and φ∈L2(Ω) be a nonnegative function satisfying the following:
∫Ωφ≥α and ∫Ωφ2≤β. |
Then,
|{x∈Ω|φ(x)≥α22|Ω|}|≥α24β. |
Lemma 2.4. [7] There exists a constant C(γ)>0 for any γ>0 such that the inequality
∫Ωψ2≤C(γ)∫Ω|∇ψ|2 |
holds for every ψ∈W1,2(Ω) satisfying
|{x∈Ω|ψ=0}|≥γ. |
Lemma 2.5. [7] Suppose ψ>0 belong to C1(ˉΩ) and |{x∈Ω|ψ≥ξ}|≥γ for every ξ>0,γ>0. Then,
∫Ωlnψ≥|Ω|lnξ−√C(γ)|Ω|⋅∫Ω|∇ψ|2ψ2 |
with the C(γ) taken from Lemma 2.4.
Next, we employ a generalized form of Gagliardo-Nirenberg inequality [37].
Lemma 2.6. [37] Consider a bounded domain Ω⊂Rn(n≥1) with a smooth boundary. Let 0<r<p<∞, and let λ∈(0,1) be determined by the following identity:
−np=(1−n2)λ−nr(1−λ); |
then, there exists a positive constant C such that
‖ϕ‖Lp(Ω)≤C(‖∇ϕ‖λL2(Ω)‖ϕ‖1−λLr(Ω)+‖ϕ‖Lr(Ω)) |
for all ϕ∈W1,2(Ω)∩Lr(Ω).
First, we establish the L1 boundedness of the solution.
Lemma 3.1. Let n≥1, for all t∈(0,Tmax); the solution of system (1.5) satisfies the following properties:
∫Ωu(⋅,t)≤m1:=max{∫Ωu0,|Ω|}, | (3.1) |
∫Ωv(⋅,t)≤m2:=max{∫Ωv0,|Ω|}, | (3.2) |
∫Ωw(⋅,t)≤m3:=m1+m2+∫Ωw0 | (3.3) |
and
∫Ωz(⋅,t)≤m4:=m3+∫Ωz0. | (3.4) |
Proof. By integrating the first equation of (1.5), it follows that
ddt∫Ωu=μ1∫Ωu−μ1∫Ωu2−μ1a1∫Ωuv. | (3.5) |
Invoking the Hölder inequality, we derive the following:
ddt∫Ωu≤μ1∫Ωu−μ1∫Ωu2≤μ1∫Ωu−μ1|Ω|(∫Ωu)2, | (3.6) |
for all t∈(0,Tmax). Next, by utilizing an ODE comparison argument, we can deduce (3.1). Similarly, employing the same method, we can easily obtain (3.2)–(3.4).
Subsequently, we present the following result concerning the size of specific time sets where ∫Ωu2 is large.
Lemma 3.2. Suppose t0≥0, L>0 and T∈(0,Tmax−t0)>0. Then,
|{t∈(t0,t0+T)|∫Ωu2>L}|≤μ1m1T+m1μ1L | (3.7) |
and
∫t0+Tt0∫Ωu2+∫t0+Tt0∫Ωv2≤μ1m1T+m1μ1+μ2m2T+m2μ2 | (3.8) |
where m1,m2 are given by (3.1) and (3.2), respectively.
Proof. Integrating in time for (3.6), we derive the following:
μ1∫t0+Tt0∫Ωu2≤μ1∫t0+Tt0∫Ωu+∫Ωu(⋅,t0)−∫Ωu(⋅,t0+T)≤μ1∫t0+Tt0∫Ωu+∫Ωu(⋅,t0)≤μ1m1T+m1. | (3.9) |
Similarly, we can obtain the following:
μ2∫t0+Tt0∫Ωv2≤μ2m2T+m2. |
Setting
G1:={t∈(t0,t0+T)|∫Ωu2>L}, |
we have
∫t0+Tt0∫Ωu2≥∫G1∫Ωu2≥L|G1|. |
In view of (3.9), this readily yields (3.7).
In the following process of establishing differential inequalities, we will often use the boundedness of ∫Ωw2 and ∫Ωz2.
Lemma 3.3. There exist two constants Cz,Cw>0, for all t∈(0,Tmax); the components w and z of the solution satisfy
∫Ωw2≤Cw | (3.10) |
and
∫Ωz2≤Cz. | (3.11) |
Proof. Invoking the Poincaré inequality, we have the following:
∫Ωh2≤c1∫Ω|∇h|2+1|Ω|(∫Ωh)2 |
for all h∈W1,2(Ω), where c1>0. Therefore, by utilizing (3.3), we derive the following:
∫Ω|∇w|2≥1c1∫Ωw2−1c1|Ω|m23. | (3.12) |
By multiplying the third equation of (1.5) by w, combining Young's inequality, and integrating by parts, we have the following:
12ddt∫Ωw2=−∫Ω|∇w|2+∫Ω(−w2+wu+wv)≤−∫Ω|∇w|2+12∫Ω(u2+v2). |
By substituting (3.12) into the inequality mentioned above, we derive the following:
ddt∫Ωw2+2c1∫Ωw2≤∫Ω(u2+v2)+2m23c1|Ω|. |
The inequality (3.8) yields the following:
∫t+1t{∫Ωu2(⋅,s)+∫Ωv2(⋅,s)+2m23c1|Ω|}≤μ1m1+m1μ1+μ2m2+m2μ2+2m23c1|Ω|. |
Invoking Lemma 2.2, it follows that
∫Ωw2≤max{∫Ωw20+c2,c1c22+2c2}:=Cw. |
for all t>0, where c2>0. This concludes the proof of (3.10).
Next, by testing the fourth equation of (1.5) with 2z and applying Young's inequality, we can deduce the following:
ddt∫Ωz2=2∫Ω(zΔz−z2+zw)≤−2∫Ω|∇z|2−2∫Ωz2+∫Ωz2+14∫Ωw2=−∫Ωz2+Cw4. |
Furthermore, we derive the following:
∫Ωz2≤max{∫Ωz20,Cw4}:=Cz. |
Lemma 3.4. For a sufficiently small ϵ>0, for all t∈(0,Tmax), the solution of (1.5) satisfies the following:
ddt∫Ωlnu≥4ϵ1+4ϵ∫Ω|∇u|2u2−1+4ϵ4χ21∫Ω|∇z|2z2k+μ1|Ω|−μ1∫Ωu−a1μ1m2. | (3.13) |
Proof. By multiplying the first equation of (1.5) by 1u and applying integration by parts, we derive the following:
ddt∫Ωlnu=∫Ω∇(D(u)∇u)u−χ1∫Ω∇(uzk∇z)u+∫Ωμ1(1−u−a1v)=∫ΩD(u)|∇u|2u2−χ1∫Ω∇uu⋅∇zzk+μ1|Ω|−μ1∫Ωu−a1μ1∫Ωv. | (3.14) |
Employing Young's inequality, we have the following:
∫Ω∇uu⋅∇zzk≤1(1+4ϵ)χ1∫Ω|∇u|2u2+1+4ϵ4χ1∫Ω|∇z|2z2k. |
Combining (1.7) and (3.14), we derive the following:
ddt∫Ωlnu≥4ϵ1+4ϵ∫Ω|∇u|2u2−1+4ϵ4χ21∫Ω|∇z|2z2k+μ1|Ω|−μ1∫Ωu−a1μ1m2. |
Lemma 3.5. For all t∈(0,Tmax), it follows that
ddt∫Ωz2−2k=−2(1−k)(1−2k)∫Ω|∇z|2z2k−2(1−k)∫Ωz2−2k+2(1−k)∫Ωz1−2kw. | (3.15) |
Proof. By multiplying the fourth equation of (1.5) by z1−2k and integrating by parts, we readily derive equation (3.15).
Next, we combine Lemmas 3.4 and 3.5, along with the boundedness of ∫Ωz2 and ∫Ωw2 to establish the Lyapunov functional, which directly affects the estimation of the integral ∫Ωu time set and provide the fundamental groundwork for proving the integral's lower bound. Based on (3.15), the range of k influences the estimation of ddt∫Ωz2−2k. Therefore, we will examine this issue in three distinct cases: k∈(0,12), k∈(12,1) and k=1.
Case 1: In this case, we consider the scenario where k∈(0,12). This allows us to obtain 1−2k∈(0,1),1−k∈(12,1); then,
ddt∫Ωz2−2k≤−2(1−k)(1−2k)∫Ω|∇z|2z2k+2(1−k)∫Ωz1−2kw. | (3.16) |
Lemma 3.6. Let n≥1, χ1>0, 0<a1<|Ω|m2, μ1=μ1(χ1,k,a1,Ω,u0,v0,w0,z0) be a large enough positive constant, k∈(0,12) and D(u) satisfy (1.6), (1.7). Then, for all t∈(0,Tmax), the solution of (1.5) satisfies the following:
ddt{∫Ωlnu−B∫Ωz2−2k}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+A|Ω|, | (3.17) |
where A,B>0.
Proof. From (3.13) and (3.16), we can derive the following:
ddt{∫Ωlnu−(1+4ϵ)χ218(1−k)(1−2k)∫Ωz2−2k}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu−(1+4ϵ)χ214(1−2k)∫Ωz1−2kw+μ1|Ω|−a1μ1m2. |
Invoking Young's inequality, (3.10), (3.11) and considering the condition k∈(0,12), we derive the following:
(1+4ϵ)χ214(1−2k)∫Ωz1−2kw≤(1+4ϵ)χ214(1−2k){(1−2k)|Ω|2Cz∫Ωz2+(1+2k)Cz2|Ω|∫Ωw21+2k}≤(1+4ϵ)χ218|Ω|+(1+4ϵ)χ21(1+2k)Cz8(1−2k)|Ω|{(1−2k)|Ω|(1+2k)Cz|Ω|+2kCz(1+2k)(1−2k)∫Ωw2}≤(1+4ϵ)χ214|Ω|(1+χ21C2zCw(1−2k)2|Ω|2). |
Combining the above inequality, we can obtain the following:
ddt{∫Ωlnu−(1+4ϵ)χ218(1−k)(1−2k)∫Ωz2−2k}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+[(1−a1m2|Ω|)μ1−(1+4ϵ)χ214|Ω|(1+χ21C2zCw(1−2k)2|Ω|2)]. |
Since μ1>0, 0<a1<|Ω|m2, we fix ϵ>0 such that it is sufficiently small to satisfy the following:
A:=(1−a1m2|Ω|)μ1−(1+4ϵ)χ214|Ω|(1+χ21C2zCw(1−2k)2|Ω|2)>0. |
Furthermore, let B:=(1+4ϵ)χ218(1−k)(1−2k)>0, it can readily yield that (3.17) holds.
Lemma 3.7. Let n≥1, χ1>0, 0<a1<|Ω|m2, μ1=μ1(χ1,k,a1,Ω,u0,v0,w0,z0) be a large enough positive constant, k∈(0,12) and D(u) satisfy (1.6), (1.7). Suppose t0≥0, K0≥0 satisfies the following:
∫Ωlnu(⋅,t0)≥−K0 | (3.18) |
and for any T∈(0,Tmax−t0)>0 fulfills the following:
T≥K0+m1+BC112A|Ω| | (3.19) |
with m1 and C1 given by (3.1) and (3.25); we can derive
∫t0+Tt0∫Ωu dxdt≥AT|Ω|2μ1, | (3.20) |
along with
|{t∈(t0,t0+T)|∫Ωu(⋅,t)≥ε}|≥εTm1, | (3.21) |
where ε is defined as
ε:=min{A|Ω|4μ1,m1}. | (3.22) |
Proof. By integrating (3.16) in time from t0 to t0+T, we can derive the following:
∫Ωlnu(⋅,t0+T)−∫Ωlnu(⋅,t0)−B∫Ωz2−2k(⋅,t0+T)+B∫Ωz2−2k(⋅,t0)≥4ϵ1+4ϵ∫t0+Tt0∫Ω|∇u|2u2−μ1∫t0+Tt0∫Ωu+AT|Ω|. |
Since u,z are positive and (3.18), we deduce the following:
μ1∫t0+Tt0∫Ωu≥∫Ωlnu(⋅,t0)−∫Ωlnu(⋅,t0+T)+B∫Ωz2−2k(⋅,t0+T)−B∫Ωz2−2k(⋅,t0)+AT|Ω|≥−∫Ωlnu(⋅,t0+T)−B∫Ωz2−2k(⋅,t0)−K0+AT|Ω|. | (3.23) |
Considering that lnξ<ξ for ξ>0, we can conclude that
−∫Ωlnu(⋅,t0+T)≥−∫Ωu(⋅,t0+T)≥−m1. | (3.24) |
By utilizing Lemma 3.3 and Young's inequality, we can derive the following:
∫Ωz2−2k≤∫Ωz2+|Ω|(1−k)1−kkk:=C1, | (3.25) |
where C1>0.
Therefore, invoking (3.19), (3.24) and (3.25), the inequality (3.23) can be restated as follows
∫t0+Tt0∫Ωu≥1μ1(A|Ω|T−K0−m1−BC1)≥A|Ω|T2μ1, |
which yields (3.20). Next, let
G2:={t∈(t0,t0+T)|∫Ωu(⋅,t)≥ε}; |
it follows that
∫t0+Tt0∫Ωu=∫G2∫Ωu+∫(t0,t0+T)∖G2∫Ωu≤m1|G2|+εT. |
Thanks to (3.20), we derive the following:
|G2|≥A|Ω|T2μ1m1−εTm1. |
The proof of Lemma 3.7 has been completed.
For another application of (3.17), we can get the size of the time set where ∫Ω|∇u|2u2 is big enough.
Lemma 3.8. Let n≥1, T∈(0,Tmax−t0)>0, χ1>0, 0<a1<|Ω|m2, μ1=μ1(χ1,k,a1,Ω,u0,v0,w0,z0) be a large enough positive constant, k∈(0,12) and D(u) satisfy (1.6), (1.7), suppose that (3.13) is true. Then, we have the following:
|{t∈(t0,t0+T)|∫Ω|∇u|2u2>M}|≤(1+4ϵ)(m1+K0+BC1+μ1m1T)4ϵM | (3.26) |
where M>0.
Proof. Integrating (3.16) over t∈(t0,t0+T), we derive the following:
4ϵ1+4ϵ∫t0+Tt0∫Ω|∇u|2u2≤∫Ωlnu(⋅,t0+T)−∫Ωlnu(⋅,t0)−B∫Ωz2−2k(⋅,t0+T)+B∫Ωz2−2k(⋅,t0)+μ1∫t0+Tt0∫Ωu−AT|Ω|. |
Hence, by utilizing (3.18), (3.25) and positive of z, we have the following:
4ϵ1+4ϵ∫t0+Tt0∫Ω|∇u|2u2≤m1+K0+BC1+μ1m1T. | (3.27) |
We define
G3:={t∈(t0,t0+T)|∫Ω|∇u|2u2>M}; |
then,
4ϵ1+4ϵ∫t0+Tt0∫Ω|∇u|2u2≥4ϵ1+4ϵ∫G3∫Ω|∇u|2u2≥4ϵ1+4ϵM|G3|. |
Combining with (3.27), it yields the following:
|G3|≤m1+K0+BC1+μ1m1T4ϵ1+4ϵM. |
By applying the aforementioned lemmas, we can combine the time sets concerning ∫Ωu, ∫Ωu2, and ∫Ω|∇u|2u2 to obtain an upper bound on ∫Ωlnu. The following proof process is based on the method described in [7].
Lemma 3.9. Let n≥1, T∈(0,Tmax−t0)>0, χ1>0, 0<a1<|Ω|m2, μ1=μ1(χ1,k,a1,Ω,u0,v0,w0,z0) be a large enough positive constant, k∈(0,12) and D(u) satisfy (1.6), (1.7), if (u0,v0,w0,z0) satisfy initial condition and
∫Ωu0≤m1, ∫Ωw20≤Cw, ∫Ωz20≤Cz, ∫Ωlnu0≥−K0. | (3.28) |
Then, there exist constants K1 and a sequence (ti)i∈N⊂[0,∞) fulfills ti→∞ as i→∞ and ti<ti+1<ti+T along with
∫Ωlnu(⋅,ti)≥−K1 for all i∈N. | (3.29) |
Proof. Let
ε:=min{A|Ω|4μ1,m1} | (3.30) |
be defined in Lemma 3.7; we choose L>0, M>0 such that
m1L<ε4m1 | (3.31) |
and
(1+4ϵ)μ1m1ϵM<ε2m1. | (3.32) |
Furthermore, we can choose
γ:=ε24L | (3.33) |
and
ξ:=ε2|Ω|. | (3.34) |
To demonstrate that (3.29) can be achieved through a suitably selected sequence (ti)i∈N⊂[0,∞) fulfills ti→∞ as i→∞, we define ti inductively. Initially, let t1:=0 and for every i≥1 and with the assumption that t1,⋯,ti possess the following property
∫Ωlnu(⋅,tk)≥−K1 | (3.35) |
for all k∈{1,⋯,i}, where
K1:=max{K0,−|Ω|lnξ+√C(γ)|Ω|M}, | (3.36) |
then one can find ti+1∈(ti+εT4m1,ti+T) such that (3.35) holds for j=i+1.
Let
Q1:={t∈(ti,ti+T)|∫Ωu(⋅,t)≥ε}, |
Q2:={t∈(ti,ti+T)|∫Ωu2(⋅,t)≤L}, |
Q3:={t∈(ti,ti+T)|∫Ω|∇u(⋅,t)|2u2(⋅,t)≤M}, |
we need to ensure that
|Q1∩Q2∩Q3∩(ti+εT4m1,ti+T)|>0. | (3.37) |
Prior to proving it, we fix T>0 enough such that it satisfies
m1μ1L<εT4m1, | (3.38) |
(1+4ϵ)(m1+K1+BC1)ϵM<εT2m1, | (3.39) |
and
T>K1+m1+BC112A|Ω|. | (3.40) |
Clearly, we can set K0:=K1 and t0:=ti. By (3.35) and (3.40), we can infer (3.18) and (3.19). Furthermore, invoking Lemma 3.7, we can directly derive the following:
|Q1|>εTm1, | (3.41) |
where ε is defined in (3.30). Subsequently, combining (3.28) and (3.31), we can employ Lemma 3.2 with t0:=ti to obtain the following:
|Q2|=T−|{t∈(ti,ti+T)|∫Ωu2(⋅,t)>L}|≥T−μ1Tm1+m1μ1L=(1−m1L)T−m1μ1L>(1−ε2m1)T, | (3.42) |
where 1−ε2m1>0. Notably, due to (3.22), we have ε≤m1. With this observation in mind, considering (3.29) (3.32) (3.39), and using Lemma 3.8, we obtain the following:
|Q3|=T−|{t∈(ti,ti+T)|∫Ω|∇u(⋅,t)|2u2(⋅,t)>M}|≥T−(1+4ϵ)μ1Tm14ϵM−(1+4ϵ)(m1+K0+BC1)4ϵM>T−εT8m1−εT8m1>(1−ε4m1)T. |
Given (3.41) and (3.42), it yields the following:
|Q1∩Q2|>εT2m1. |
Therefore, it follows that
|Q1∩Q2∩Q3|>εT2m1+(1−ε4m1)T−T=εT4m1, |
which clearly yields (3.37).
Utilizing (3.37) and setting t0=0, we can find a t1>t0+εT4m1 such that t1∈Q1∩Q2∩Q3. Subsequently, by Lemmas 2.3 and 2.5 and (3.36), we derive the following:
∫Ωlnu(⋅,t1)≥−K1, |
where Lemma 2.5 is applied according to the definitions (3.33) and (3.34) of γ and ξ.
Similarly, we can also find t2>t1+εT4m1 such that
∫Ωlnu(⋅,t2)≥−K1. |
Therefore, we can find a sequence {ti} satisfying
ti+T>ti+1>ti+εT4m1 |
and
∫Ωlnu(⋅,ti)≥−K1, |
without a loss of the fact ti→∞ as i→∞.
Case 2: In this case, we consider the scenario where k∈(12,1). This allows us to obtain the following:
1−2k∈(−1,0), 1−k∈(0,12); |
then,
ddt∫Ωz2−2k≥2(1−k)(2k−1)∫Ω|∇z|2z2k−2(1−k)∫Ωz2−2k. | (3.43) |
Lemma 3.10. Let n≥1, χ1>0, 0<a1<|Ω|m2, μ1=μ1(χ1,k,a1,Ω,u0,v0,w0,z0) be a large enough positive constant, k∈(12,1) and D(u) satisfy (1.6), (1.7). Then, for all t∈(0,Tmax), the solution of (1.5) satisfies the following:
ddt{∫Ωlnu+D∫Ωz2−2k}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+E|Ω|, | (3.44) |
where D,E>0.
Proof. We can use Young's inequality to obtain the following:
∫Ωz2−2k≤2(1−k)∫Ωz+(2k−1)|Ω|≤2(1−k)m4+(2k−1)|Ω|. |
By combining (3.13) and (3.43), we can obtain the following:
ddt{∫Ωlnu+(1+4ϵ)χ218(1−k)(2k−1)∫Ωz2−2k}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+μ1|Ω|(1+4ϵ)χ214|Ω|−2(1+4ϵ)χ21(1−k)m44(2k−1)−a1μ1m2≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+[(1−a1m2|Ω|)μ1−(1+4ϵ)χ214(1+m42k−1)]|Ω|. |
Since μ1>0 is large enough and 0<a1<|Ω|m2, we fix ϵ>0 sufficiently small such that
E:=(1−a1m2|Ω|)μ1−(1+4ϵ)χ214(1+m42k−1)>0. |
Thus, let D:=(1+4ϵ)χ218(1−k)(2k−1)>0, which entails (3.44).
By using a similar method in proving Lemmas 3.7–3.9, we can obtain a lower bounded estimate of ∫Ωlnu(⋅,ti), ti→∞ as i→∞. We have omitted some details here.
Case 3: Finally, we consider the case of k=1, it follow from (3.13) that
ddt∫Ωlnu≥4ϵ1+4ϵ∫Ω|∇u|2u2−1+4ϵ4χ21∫Ω|∇z|2z2+μ1|Ω|−μ1∫Ωu−a1μ1m2. |
Utilizing the fourth equation of (1.5), we can derive the following:
ddt∫Ωlnz=∫Ω|∇z|2z2−|Ω|+wz≥∫Ω|∇z|2z2−|Ω|, |
for all t∈(0,Tmax). We can combine the above two inequality to obtain the following:
ddt{∫Ωlnu+1+4ϵ4χ21∫Ωlnz}≥4ϵ1+4ϵ∫Ω|∇u|2u2−μ1∫Ωu+F|Ω|, | (3.45) |
where F:=μ1−a1μ1m2|Ω|−1+4ϵ4χ21>0. Because we chose a large enough μ1>0 and 0<a1<|Ω|m2, we can also select a small enough value for ϵ>0 in order to achieve the following:
μ1>(1+4ϵ)χ21|Ω|4(|Ω|−a1m2). |
Then, we will structure the lower bound for ∫Ωlnz.
Lemma 3.11. There exists a constant S>0, for all t∈(0,Tmax); the component z of the solution satisfies the following:
∫Ωlnz≥−S. |
Proof. By employing a similar methodology to that presented in Lemma 2.3 of [33], we can easily establish the validity of Lemma 3.11. However, in the interest of brevity, we refrain from providing the details here.
By applying the method of discussing k∈(0,12), combining with the reconstructed functional (3.45) and Lemma 3.11, we can also find a lower bound for ∫Ωlnu(⋅,ti), ti→∞ as i→∞.
Proof of Theorem 1.1 We infer from (3.5) that
ddt∫Ωu(⋅,t)≤μ1∫Ωu(⋅,t). |
By applying Lemmatas 3.9–3.11, it indicates that
∫Ωu(⋅,t)≥∫Ωu(⋅,ti)e−μ1(ti−t)≥εe−μ1(ti−t), | (3.46) |
where t∈[0,ti) and i∈N. Consequently, we directly obtain ∫Ωu(⋅,t)≥εe−μ1t1. We observe that the inequality ti+1<ti+T holds for all i≥1 and large values of t. Hence, we can denote (3.46) as ∫Ωu(⋅,t)≥εe−2μ1T for all t∈[ti,ti+1). Notably, without a loss of generality, ti→∞ as i→∞, this implies that it possesses a lower bound mu:=min{εe−μ1t1,εe−2μ1T} for ∫Ωu. Through constructing some other Lyapunov functionals, we can then utilize the same methodology to determine lower the bound mv for ∫Ωv.
Assuming that Ω⊂Rn(n≥2) is a bounded convex domain with a smooth boundary, for the sake of convenience, we define θ:=min{θ1,θ2}. First, we establish a lower bound for z based on Theorem 1.1; other approach details can be also found in [40].
Lemma 4.1. Assuming that the conditions in Theorem 1.1 hold and t∈[0,Tmax); then, we can find a constant δ>0 independent of t such that
infx∈Ωz(x,t)≥δ>0. | (4.1) |
Proof. By taking the positivity of u and applying the comparison principle, we can infer from the fourth equation of system (1.5) that
z(x,t)≥δ(t):=infy∈Ωz0(y)e−t. | (4.2) |
Let us fix δ1=12infx∈Ωz0(x); then, there exists a t0>0 such that z(x,t)>δ1 for t∈[0,t0]. To conclude the proof, it is sufficient to show that t∈[t0,Tmax). By referencing the established result of Lemma 3.1 in [38] and for all nonnegative φ∈C0(ˉΩ), we can derive the following:
etΔφ≥1(4πt)n2e−(diamΩ)24t∫Ωφdx. | (4.3) |
By virtue of the variation of constants formula and Theorem 1.1, we derive the following:
w(x,t)=et(Δ−1)w0+∫t0e(t−s)(Δ−1)(u(x,t)+v(x,t))ds≥∫t01(4π(t−s))n2e−(t−s)−(diamΩ)24(t−s)∫Ω(u(x,t)+v(x,t))dxds≥(mu+mv)∫t001(4πs)n2e−s−(diamΩ)24sds=:c>0. | (4.4) |
Reusing the variation of constants formula to z and (4.4), we deduce the following:
z(x,t)=et(Δ−1)z0+∫t0e(t−s)(Δ−1)w(x,t)ds≥∫t01(4π(t−s))n2e−(t−s)−(diamΩ)24(t−s)∫Ωw(x,t)dxds≥c|Ω|∫t001(4πs)n2e−s−(diamΩ)24sds=:δ2>0. |
Now, we can set δ:=min{δ1,δ2} to complete the proof.
In order to more conveniently utilize the Gagliardo-Nirenberg inequality in Lemma 4.7, we need to select some parameters in advance. For any p>1 and q>1, we define
α1=2(p+1)1+θ, | (4.5) |
α2=2(p+1)(q−1)p−1, | (4.6) |
λi=q−qαiq+1n−12 | (4.7) |
and
fi=αiqλi=αi−1q+1n−12 | (4.8) |
for i=1,2.
Lemma 4.2. Let n≥2; then, for θ>1−2n+1 and a sufficiently large p>1, there exists a q>1 such that
λi∈(0,1) and fi<2, for i=1,2. | (4.9) |
Proof. First, a simple calculation reveals that the first inequality in (4.9) is equivalent to the following:
0<q−qαi<q+1n−12, for i=1,2. | (4.10) |
Regarding the above inequality, we can obtain αi>1 and q>αi2−1n. Additionally, fi<2 is equivalent to αi−1<2(q+1n−12), which means that q>αi2−1n. Therefore, we can conclude that (4.9) holds if αi>1 and q>αi2−1n. Moreover, we need to make sure
p>1+θ2−1, q>p−12(p+1)+1, p+11+θ−1n<q<p+12+p−12n. |
Thus, the existence of q is dependent on p+11+θ<p+12+p−12n+1n, which can be easily derived thanks to θ>1−2n+1. Thus, when p is large enough, we can choose a q such that the inequality (4.10) makes sense. As a result, we can conclude that (4.9) is valid.
Lemma 4.3. Let n≥2; there is a constant C1>0 such that
‖∇z(⋅,t)‖L1(Ω)≤C1, for all t∈(0,Tmax). | (4.11) |
Proof. By applying integration by parts and Young's inequality, we can obtain the following result when testing the fourth equation of (1.5) with −2Δz:
ddt∫Ω|∇z|2=−2∫Ω(|Δz|2+|∇z|2+wΔz)≤−2∫Ω|Δz|2−2∫Ω|∇z|2+2∫Ω|Δz|2+12∫Ωw2=−2∫Ω|∇z|2+12∫Ωw2. |
By applying (3.10) and the ODE comparison principle, we have ∫Ω|∇z|2≤c1, where c1>0 is a constant. By utilizing Young's inequality once more, we obtain ∫Ω|∇z|≤∫Ω|∇z|2+14|Ω|≤c2, where c2>0 is a constant. Therefore, we can easily derive (4.11).
Lemma 4.4. Assuming that the conditions stated in Theorem 1.1 hold, we can conclude that for any p>1, for all t∈(0,Tmax), there exist positive constants C2,C3 that are independent of t, such that
ddt∫Ωup+∫Ωup≤−μ1∫Ωup+1+C2∫Ω|∇z|2(p+1)1+θ+C2 | (4.12) |
and
ddt∫Ωvp+∫Ωvp≤−μ2∫Ωvp+1+C3∫Ω|∇z|2(p+1)1+θ+C3. | (4.13) |
Proof. By multiplying the first equation of (1.5) by up−1 and performing integration by parts, we derive the following:
1pddt∫Ωup=∫Ωup−1∇(D(u)∇u)−χ1∫Ωup−1∇(uzk∇z)+μ1∫Ωup(1−u−a1v)≤−(p−1)∫Ωup−2+θ|∇u|2+χ1(p−1)δk∫Ωup−1∇u⋅∇z+μ1∫Ωup−μ1∫Ωup+1−μ1a1∫Ωupv, | (4.14) |
where δ is from Lemma 4.1. By applying Young's inequality twice, we deduce the following:
∫Ωup−1∇u⋅∇z≤δk2χ1∫Ωup−2+θ|∇u|2+μ1δk2χ1p(p−1)up+1+c3∫Ω|∇z|2(p+1)1+θ, | (4.15) |
where c3>0 is a constant. Invoking the Young's inequality once more, we have the following:
(μ1+1p)∫Ωup≤(2p−3)μ12p∫Ωup+1+c4, | (4.16) |
where c4>0. Together (4.14)–(4.16), we can derive (4.12). Furthermore, a similar argument used to obtain the inequality in (4.12) implies that (4.13) also holds.
To obtain the first term on the right side of inequalities (4.12) and (4.13), we introduce a differential inequality about ∫Ωwp+1 for any p>1.
Lemma 4.5. Assuming that the conditions stated in Theorem 1.1 hold, for all t∈(0,Tmax), the solution of the system (1.5) satisfies the following:
ddt∫Ωwp+1+∫Ωwp+1≤2p∫Ωup+1+2p∫Ωvp+1. | (4.17) |
Proof. By multiplying the third equation in (1.5) with (p+1)wp and Young's inequality, we have the following:
ddt∫Ωwp+1=−p(p+1)∫Ωwp−1|∇w|2−(p+1)∫Ωwp+1+(p+1)∫Ωwpu+(p+1)∫Ωwpv≤−∫Ωwp+1+2p∫Ωup+1+2p∫Ωvp+1, |
which implies (4.17).
To hand the second term on the right side of inequalities (4.12) and (4.13), we will outline the properties of the component solution z of system (1.5).
Lemma 4.6. Assuming that the conditions stated in Theorem 1.1 hold, for all t∈(0,Tmax), for any q>1, there exists C4>0 independent of t such that
ddt∫Ω|∇z|2q+2q∫Ω|∇z|2q≤−2(q−1)q∫Ω|∇|∇z|q|2+12∫Ωwp+1+C4∫Ω|∇z|2(p+1)(q−1)p−1. | (4.18) |
Proof. The result is a well-established inequality, which can be found in Lemma 4.2 of [41]. Therefore, we have omitted the proof process, and interested readers are encouraged to refer to the original literature and the references therein for further details.
By invoking the above four lemmas, we can provide a bound for Lp norm of u for any p>1.
Lemma 4.7. Let n≥2. Assuming that the conditions stated in Theorem 1.1 hold, and considering θ>1−2n+1, then for all t∈(0,Tmax) and any p>1, we can find a constant C5>0 independent of t such that
∫Ωup≤C5 and ∫Ωvp≤C5. | (4.19) |
Proof. Combing (4.12), (4.13), (4.17) and (4.18), we can derive the following:
ddt{2pμ1∫Ωup+2pμ2∫Ωvp+∫Ωwp+1+∫Ω|∇z|2q}+2pμ1∫Ωup+2pμ2∫Ωvp+12∫Ωwp+1+2q∫Ω|∇z|2q+2(q−1)q∫Ω|∇|∇z|q|2≤c5∫Ω|∇z|α1+c6∫Ω|∇z|α2+c7, | (4.20) |
where c5,c6,c7>0 and α1,α2 are defined in (4.5) and (4.6). In view of the Gagliardo-Nirenberg inequality, there exists constants c8,c9>0 such that
c5∫Ω|∇z|α1=c5‖|∇z|q‖α1qLα1q(Ω)≤c8‖∇|∇z|q‖λ1α1qL2(Ω)‖|∇z|q‖α1q(1−λ1)L1q(Ω)+c8‖|∇z|q‖α1qL1q(Ω), |
c6∫Ω|∇z|α2=c6‖|∇z|q‖α2qLα2q(Ω)≤c9‖∇|∇z|q‖λ2α2qL2(Ω)‖|∇z|q‖α2q(1−λ2)L1q(Ω)+c9‖|∇z|q‖α2qL1q(Ω), |
where λi(i=1,2) are defined in (4.7) and λi∈(0,1) from Lemma 4.2. Thus, thanks to ‖|∇z|q‖L1q(Ω)=‖∇z‖qL1(Ω) and applying Young's inequality, we can infer from Lemma 4.3 that
c5∫Ω|∇z|α1≤q−1q∫Ω|∇|∇z|q|2+c10, | (4.21) |
c6∫Ω|∇z|α2≤q−1q∫Ω|∇|∇z|q|2+c11, | (4.22) |
where c10,c11>0. Substituting (4.21), (4.22) into (4.20), we obtain f′(t)+c12f(t)≤c13, where f(t):=2pμ1∫Ωup+2pμ2∫Ωvp+∫Ωwp+1+∫Ω|∇z|2q, c12:=min{12,2q}, c13>0.
Therefore, employing the standard ODE comparison principle, we can infer that (4.19) is valid.
Proof of Theorem 1.2 Due to (4.19) and Lemma 4.1 in [34], for all t∈(0,Tmax), we can readily deduce that for all σ>1,
‖w(⋅,t)‖W1,σ(Ω)≤C6, |
where C6>0 is a constant.
By applying some parabolic regularity and utilizing Lemma 4.3, for all t∈(0,Tmax), we obtain the following result:
‖w(⋅,t)‖W1,∞(Ω)+‖z(⋅,t)‖W1,∞(Ω)≤C7, |
where C7>0. Taking advantage of a standard Alikakos-Moser iteration [39] and Lemma 4.6, for all t∈(0,Tmax), there exists a constant C8>0 such that
‖u(⋅,t)‖L∞(Ω)+‖v(⋅,t)‖L∞(Ω)≤C8. |
By combining with Lemma 2.1, we can establish the validity of Theorem 1.2.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare there is no conflict of interest.
This research was supported by Guizhou Provincial Basic Research Program (Natural Science) (Nos. QKHJC-ZK[2023]YB140, QKHJC-ZK[2021]YB317), and by the Natural Science Research Project of Department of Education of Guizhou Province (Nos. QJJ2023012, QJJ2023061, QJJ2023062).
[1] |
T. Hillen, K. J. Painter, A user's guide to PDE models for chemotaxis, J. Math. Biol., 58 (2009), 183–217. https://doi.org/10.1007/s00285-008-0201-3 doi: 10.1007/s00285-008-0201-3
![]() |
[2] |
E. F. Keller, L. A. 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
![]() |
[3] |
D. Horstmann, G. Wang, Blow-up in a chemotaxis model without symmetry assumptions, Eur. J. Appl. Math., 12 (2001), 159–177. https://doi.org/10.1017/S0956792501004363 doi: 10.1017/S0956792501004363
![]() |
[4] |
M. Winkler, Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source, Commun. Part. Differ. Equation, 35 (2010), 1516–1537. https://doi.org/10.1080/03605300903473426 doi: 10.1080/03605300903473426
![]() |
[5] |
J. Lankeit, Eventual smoothness and asymptotics in a three-dimensional chemotaxis system with logistic source, J. Differ. Equation, 258 (2015), 1158–1191. https://doi.org/10.1016/j.jde.2014.10.016 doi: 10.1016/j.jde.2014.10.016
![]() |
[6] |
T. Xiang, How strong a logistic damping can prevent blow-up for the minimal Keller-Segel chemotaxis system?, J. Math. Anal. Appl., 459 (2018), 1172–1200. https://doi.org/10.1016/j.jmaa.2017.11.022 doi: 10.1016/j.jmaa.2017.11.022
![]() |
[7] |
Y. S. Tao, M. Winkler, Persistence of mass in a chemotaxis system with logistic source, J. Differ. Equation, 259 (2015), 6142–6161. https://doi.org/10.1016/j.jde.2015.07.019 doi: 10.1016/j.jde.2015.07.019
![]() |
[8] |
M. Ding, W. Wang, S. Zhou, Global existence of solutions to a fully parabolic chemotaxis system with singular sensitivity and logistic source, Nonlinear Anal. Real, 49 (2019), 286–311. https://doi.org/10.1016/j.nonrwa.2019.03.009 doi: 10.1016/j.nonrwa.2019.03.009
![]() |
[9] |
J. Lankeit, Chemotaxis can prevent thresholds on population density, Discrete Contin. Dyn. B, 20 (2015), 1499–1527. https://doi.org/10.3934/dcdsb.2015.20.1499 doi: 10.3934/dcdsb.2015.20.1499
![]() |
[10] |
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
![]() |
[11] |
K. J. Painter, T. Hillen, Spatio-temporal chaos in a chemotaxismodel, Phys. D, 240 (2011), 363–375. https://doi.org/10.1016/j.physd.2010.09.011 doi: 10.1016/j.physd.2010.09.011
![]() |
[12] |
S. Strohm, R. C. Tyson, J. A. Powell, Pattern formation in a model for mountain pine beetle dispersal: Linking model predictions to data, Bull. Math. Biol., 75 (2013), 1778–1797. https://doi.org/10.1007/s11538-013-9868-8 doi: 10.1007/s11538-013-9868-8
![]() |
[13] |
B. Hu, Y. Tao, To the exclusion of blow-up in a three-dimensional chemotaxis-growth model with indirect attractant production, Math. Mod. Meth. Appl. S, 26 (2016), 2111–2128. https://doi.org/10.1142/S0218202516400091 doi: 10.1142/S0218202516400091
![]() |
[14] |
J. Xing, P. Zheng, Y. Xiang, H. Wang, On a fully parabolic singular chemotaxis-(growth) system with indirect signal production or consumption, Z. Angew. Math. Phys., 72 (2021), 105. https://doi.org/10.1007/s00033-021-01534-6 doi: 10.1007/s00033-021-01534-6
![]() |
[15] |
C. Liu, G. Ren, B. Liu, Boundedness in a higher-dimensional singular chemotaxis-growth system with indirect signal production, Z. Angew. Math. Phys., 74 (2023), 119. https://doi.org/10.1007/s00033-023-02017-6 doi: 10.1007/s00033-023-02017-6
![]() |
[16] |
T. Black, J. Lankeit, M. Mizukami, On the weakly competitive case in a two-species chemotaxis model, IMA J. Appl. Math., 81 (2016), 860–876. https://doi.org/10.1093/imamat/hxw036 doi: 10.1093/imamat/hxw036
![]() |
[17] |
J. I. Tello, M. Winkler, Stabilization in a two-species chemotaxis system with a logistic source, Nonlinearity, 25 (2012), 1413–1425. https://doi.org/10.1088/0951-7715/25/5/1413 doi: 10.1088/0951-7715/25/5/1413
![]() |
[18] | M. Mizukami, Y. Tanaka, Finite-time blow-up in a two-species chemotaxis-competition model with single production, Arch. Math., 59 (2023), 215–222. |
[19] |
M. Mizukami, Y. Tanaka, T. Yokota, Can chemotactic effects lead to blow-up or not in two-species chemotaxis-competition models?, Z. Angew. Math. Phys., 73 (2022), 239. https://doi.org/10.1007/s00033-022-01878-7 doi: 10.1007/s00033-022-01878-7
![]() |
[20] |
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
![]() |
[21] |
X. Li, Y. Wang, On a fully parabolic chemotaxis system with Lotka–Volterra competitive kinetics, J. Math. Anal. Appl., 471 (2019), 584–598. https://doi.org/10.1016/j.jmaa.2018.10.093 doi: 10.1016/j.jmaa.2018.10.093
![]() |
[22] |
S. Qiu, C. Mu, X. Tu, Global existence and boundedness to a two-species chemotaxis-competition model with singular sensitivity, J. Evol. Equation, 21 (2021), 33–62. https://doi.org/10.1007/s00028-020-00570-5 doi: 10.1007/s00028-020-00570-5
![]() |
[23] |
S. Chen, J. Shi, Global dynamics of the diffusive Lotka–Volterra competition model with stage structure, Calculus Var. Partial Differ. Equations, 59 (2020), 33. https://doi.org/10.1007/s00526-019-1693-y doi: 10.1007/s00526-019-1693-y
![]() |
[24] |
Y. Xiang, P. Zheng, On a two-species chemotaxis-competition system with indirect signal consumption, Z. Angew. Math. Phys., 73 (2022), 50. https://doi.org/10.1007/s00033-022-01680-5 doi: 10.1007/s00033-022-01680-5
![]() |
[25] |
Y. Xiang, P. Zheng, J. Xing, Boundedness and stabilization in a two-species chemotaxis-competition system with indirect signal production, J. Math. Anal. Appl., 507 (2022), 125825. https://doi.org/10.1016/j.jmaa.2021.125825 doi: 10.1016/j.jmaa.2021.125825
![]() |
[26] |
X. Tu, C. Mu, S. Qiu, Boundedness and convergence of constant equilibria in a two-species chemotaxis-competitionsystem with loop, Nonlinear Anal., 198 (2020), 111923. https://doi.org/10.1016/j.na.2020.111923 doi: 10.1016/j.na.2020.111923
![]() |
[27] |
D. Wang, F. Zeng, M. Jiang, Global existence and boundedness of solutions to a two-species chemotaxis-competition system with singular sensitivity and indirect signal production, Z. Angew. Math. Phys., 74 (2023), 33. https://doi.org/10.1007/s00033-022-01921-7 doi: 10.1007/s00033-022-01921-7
![]() |
[28] | T. Black, Global existence and asymptotic stability in a competitive two-species chemotaxis system with two signals, Discrete Contin. Dyn. B, 22 (2017) 1253–1272. https://doi.org/10.3934/dcdsb.2017061 |
[29] |
L. Wang, C. Mu, A new result for boundedness and stabilization in a two-species chemotaxis system with two chemicals, Discrete Contin. Dyn. Syst. Ser. B, 25 (2020), 4585–4601. https://doi.org/10.3934/dcdsb.2020114 doi: 10.3934/dcdsb.2020114
![]() |
[30] |
X. Tu, C. Mu, P. Zheng, K. Lin, Global dynamics in a two-species chemotaxis-competition system with two signals, Discrete Contin. Dyn. A, 38 (2018), 3617–3636. https://doi.org/10.3934/dcds.2018156 doi: 10.3934/dcds.2018156
![]() |
[31] |
M. Winkler, Global classical solvability and generic infinite-time blow-up in quasilinear Keller–Segel systems with bounded sensitivities, J. Differ. Equaiton, 266 (2019), 8034–8066. https://doi.org/10.1016/j.jde.2018.12.019 doi: 10.1016/j.jde.2018.12.019
![]() |
[32] |
Y. Tao, M. Winkler, A chemotaxis-haptotaxis model: the roles of nonlinear diffusion and logistic source, SIAM. J. Math. Anal., 43 (2011), 685–704. https://doi.org/10.1137/100802943 doi: 10.1137/100802943
![]() |
[33] |
J. Zhao, A quasilinear parabolic-parabolic chemotaxis model with logistic source and singular sensitivity, Discrete Contin. Dyn. Syst. B, 27 (2022), 3487–3513. https://doi.org/10.3934/dcdsb.2021193 doi: 10.3934/dcdsb.2021193
![]() |
[34] |
D. Horstmann, M. Winkler, Boundedness vs. blow-up in a chemotaxis system, J. Differ. Equation, 215 (2005), 52–107. https://doi.org/10.1016/j.jde.2004.10.022 doi: 10.1016/j.jde.2004.10.022
![]() |
[35] |
E. Lankeit, J. Lankeit, Classical solutions to a logistic chemotaxis model with singular sensitivity and signal absorption, Nonlinear Anal. Real World Appl., 46 (2019), 421–445. https://doi.org/10.1016/j.nonrwa.2018.09.012 doi: 10.1016/j.nonrwa.2018.09.012
![]() |
[36] |
C. Stinner, C. Surulescu, 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
![]() |
[37] |
Y. Li, J. Lankeit, Boundedness in a chemotaxis–haptotaxis model with nonlinear diffusion, Nonlinearity, 29 (2016), 1564–1595. https://doi.org/10.1088/0951-7715/29/5/1564 doi: 10.1088/0951-7715/29/5/1564
![]() |
[38] |
T. Hillen, K. Painter, M. Winkler, Convergence of a cancer invasion model to a logistic chemotaxis model, Math. Mod. Meth. Appl. S, 23 (2013), 165–198. https://doi.org/10.1142/S0218202512500480 doi: 10.1142/S0218202512500480
![]() |
[39] |
N. Alikakos, Lp bounds of solutions of reaction-diffusion equations, Commun. Part. Differ. Equation, 4 (1979), 827–868. https://doi.org/10.1080/03605307908820113 doi: 10.1080/03605307908820113
![]() |
[40] |
T. Xiang, Finite time blow-up in the higher dimensional parabolic-elliptic-ODE minimal chemotaxis-haptotaxis system, J. Differ. Equation, 336 (2022), 44–72. https://doi.org/10.1016/j.jde.2022.07.015 doi: 10.1016/j.jde.2022.07.015
![]() |
[41] |
S. Qiu, C. Mu, L. Wang, Boundedness in the higher-dimensional quasilinear chemotaxis-growth system with indirect attractant production, Comput. Math. Appl., 75 (2018), 3213–3223. https://doi.org/10.1016/j.camwa.2018.01.042 doi: 10.1016/j.camwa.2018.01.042
![]() |
1. | S. Amalorpava Josephine, S. Karthikeyan, L. Shangerganesh, K. Yadhavan, Global Classical Solutions for a Chemotaxis System of Attraction–Repulsion With Singular Sensitivity, 2025, 0170-4214, 10.1002/mma.10806 |