In this paper, we study the following fully parabolic chemotaxis system with nonlocal growth and indirect signal production:
{ut=∇⋅(D(u)∇u)−∇⋅(u∇v)+f(u),x∈Ω,t>0,vt=Δv+w−v,x∈Ω,t>0,wt=Δw+u−w,x∈Ω,t>0,∂u∂n=∂v∂n=∂w∂n=0,x∈∂Ω,t>0,u(x,0)=u0(x),v(x,0)=v0(x),w(x,0)=w0(x),x∈Ω,
in a smooth bounded domain Ω⊂Rn(n≥3), where D(u)≥D1uγ, f(u)=u(a0−a1uσ+a2∫Ωuσdx), D1,a1,a2 are positive constants, a0,γ∈R, σ≥max{1,−γ} and a1−a2|Ω|>0. It is shown that the above system admits a globally bounded classical solution if 4n+γσ>3−σ1+σ. Furthermore, by the method of Lyapunov functionals, the global stability of steady states with convergence rates is established.
Citation: Min Jiang, Dandan Liu, Rengang Huang. Boundedness and stabilization in a quasilinear chemotaxis model with nonlocal growth term and indirect signal production[J]. Communications in Analysis and Mechanics, 2025, 17(2): 387-412. doi: 10.3934/cam.2025016
[1] | Chang-Jian Wang, Jia-Yue Zhu . Global existence and uniform boundedness to a bi-attraction chemotaxis system with nonlinear indirect signal mechanisms. Communications in Analysis and Mechanics, 2023, 15(4): 743-762. doi: 10.3934/cam.2023036 |
[2] | Shuyue Ma, Jiawei Sun, Huimin Yu . Global existence and stability of temporal periodic solution to non-isentropic compressible Euler equations with a source term. Communications in Analysis and Mechanics, 2023, 15(2): 245-266. doi: 10.3934/cam.2023013 |
[3] | Chang-Jian Wang, Jia-Yue Zhu . Analysis of global dynamics in an attraction-repulsion model with nonlinear indirect signal and logistic source. Communications in Analysis and Mechanics, 2024, 16(4): 813-835. doi: 10.3934/cam.2024035 |
[4] | Panyu Deng, Jun Zheng, Guchuan Zhu . Well-posedness and stability for a nonlinear Euler-Bernoulli beam equation. Communications in Analysis and Mechanics, 2024, 16(1): 193-216. doi: 10.3934/cam.2024009 |
[5] | Yang Liu . Global attractors for a nonlinear plate equation modeling the oscillations of suspension bridges. Communications in Analysis and Mechanics, 2023, 15(3): 436-456. doi: 10.3934/cam.2023021 |
[6] | Yang Liu, Xiao Long, Li Zhang . Long-time dynamics for a coupled system modeling the oscillations of suspension bridges. Communications in Analysis and Mechanics, 2025, 17(1): 15-40. doi: 10.3934/cam.2025002 |
[7] | Tingfu Feng, Yan Dong, Kelei Zhang, Yan Zhu . Global existence and blow-up to coupled fourth-order parabolic systems arising from modeling epitaxial thin film growth. Communications in Analysis and Mechanics, 2025, 17(1): 263-289. doi: 10.3934/cam.2025011 |
[8] | Ming Liu, Binhua Feng . Grand weighted variable Herz-Morrey spaces estimate for some operators. Communications in Analysis and Mechanics, 2025, 17(1): 290-316. doi: 10.3934/cam.2025012 |
[9] | İrem Akbulut Arık, Seda İğret Araz . Delay differential equations with fractional differential operators: Existence, uniqueness and applications to chaos. Communications in Analysis and Mechanics, 2024, 16(1): 169-192. doi: 10.3934/cam.2024008 |
[10] | Zhiyong Wang, Kai Zhao, Pengtao Li, Yu Liu . Boundedness of square functions related with fractional Schrödinger semigroups on stratified Lie groups. Communications in Analysis and Mechanics, 2023, 15(3): 410-435. doi: 10.3934/cam.2023020 |
In this paper, we study the following fully parabolic chemotaxis system with nonlocal growth and indirect signal production:
{ut=∇⋅(D(u)∇u)−∇⋅(u∇v)+f(u),x∈Ω,t>0,vt=Δv+w−v,x∈Ω,t>0,wt=Δw+u−w,x∈Ω,t>0,∂u∂n=∂v∂n=∂w∂n=0,x∈∂Ω,t>0,u(x,0)=u0(x),v(x,0)=v0(x),w(x,0)=w0(x),x∈Ω,
in a smooth bounded domain Ω⊂Rn(n≥3), where D(u)≥D1uγ, f(u)=u(a0−a1uσ+a2∫Ωuσdx), D1,a1,a2 are positive constants, a0,γ∈R, σ≥max{1,−γ} and a1−a2|Ω|>0. It is shown that the above system admits a globally bounded classical solution if 4n+γσ>3−σ1+σ. Furthermore, by the method of Lyapunov functionals, the global stability of steady states with convergence rates is established.
Chemotaxis describes the directional movement of cells or organisms in the direction of the concentration gradient of chemical signals. In order to simulate the phenomenon that cells are attracted to the high concentrations of chemical signals secreted by themselves. In 1970, Keller and Segel proposed a classical biological chemotaxis model [1] as follows
{ut=∇⋅(D(u)∇u)−∇⋅(S(u)∇v)+f(u),x∈Ω,t>0,τvt=Δv−v+u,x∈Ω,t>0,∂u∂ν=∂v∂ν=0,x∈∂Ω,t>0,u(x,0)=u0(x),τv(x,0)=τv0(x),x∈Ω, | (1.1) |
where Ω⊂Rn,τ∈{0,1},ν denotes the outward unit normal vector on ∂Ω,u(x,t) represents the density of cells and v(x,t) denotes the density of a chemical signal. ∇⋅(D(u)∇u) and −∇⋅(S(u)∇v) represent self-diffusion and cross-diffusion, respectively. The function f(u) describes cell proliferation and death. As we all know, chemotaxis research has important applications in both biology and medicine, so it has been one of the hottest research focuses in applied mathematics nowadays. The system (1.1) with τ=0 or τ=1 has been investigated extensively in the past few decades. For τ=0,D(u)≡1,S(u)=χu and f∈C1([0,∞)) is assumed to satisfy f(u)≤a−μu2 for all u≥0 with some positive constants a,μ. The solutions to (1.1) are global and bounded for arbitrarily small μ>0 with n≤2 [2], or n≥3 and μ>(n−2)χn sufficiently large[3]. When f(u)=λu−μuα with α>1,λ≥0 and μ>0, Winkler [4] introduced a concept of very weak solutions and proved global existence of such solutions for any nonnegative initial data u0∈L1(Ω) under the assumption that α>2−1n. In the case of D,S∈C2([0,∞)) and D(u)≥c0uρ,c1uq≤S(u)≤c2uq, and f(u) is a smooth function fulfilling f(0)≥0 and f(u)≤au−μu2 for all u>0 with constants a≥0 and μ≥0, Cao in [5] showed that there exists a unique global bounded classical solution. For τ=1,D(u)=1,S(u)=χu and f(u)=0 with χ>0, the system (1.1) has the global solutions with n=1 [6]; when n=2,∫Ωu0>8πχ, the solution of the system (1.1) will blow up in finite time [7], if ∫Ωu0<8πχ, the system (1.1) possesses a globally bounded classical solution [8]. In the case of n≥3, if Ω is a ball, then for arbitrarily small mass m:=∫Ωu0>0, there exists the finite-time blow-up solutions [9] with proper initial coditions. Besides, for f(u)=0,S(u)/D(u)≤K(u+ε)α with u>0,α<2N(N∈N),K>0,ε≥0, Ishida et al. [10] ruled out convexity of Ω, then established global-in-time existence and uniform-in-time boundedness of solutions. For f(u)=au−μu2,a∈R, Cao [11] used an approach based on maximal Sobolev regularity and proved that if the ratio μχ is sufficiently large, then the unique nontrivial spatially homogeneous equilibrium given by (a+μ,a+μ) is globally asymptotically stable without the restrictions τ=1 and the convexity of Ω. For the case f(u)=u(1−uγ),D(0)>0,D(u)≥K1um1 and S(u)≤K2um2,∀u≥0, Ki∈R+,mi∈R,i=1,2, Wang et al. [12] showed that the system admits global classical solutions and they are uniformly bounded in time with the parameter pair (m1,m2) lies in some specific regions and N≥2. When D(u)≃a0(u+1)−α,S(u)≃b0u(u+1)β−1 and f(u)=ru−μu1+σ, Zheng [13] showed the globally bounded classical solutions of the system (1.1) if 0<α+β<max{σ+α,2n}, or β=σ with μ sufficiently large. In addition, the signal generation may be in a nonlinear form. Zhuang [14] established that (1.1) admits a globally bounded classical solution under β<σ−1 or β=σ−1 with r=1,μ>0 sufficiently large, and the second equation replaced by vt=Δv−v+u(u+1β−1). When D(u)≥a0(u+1)−α,0≤S(u)≤b0u(u+1)β−1 with a0,b0>0 and α,β∈R, Ding [15] provided a boundedness result under α+β+γ<2n, or β+γ<1+σ, or β+γ=1+σ with μ large enough and the bounded classical solution (u,v)→((rμ)1σ,(rμ)1σ) in L∞(Ω) exponentially under the condition of b>b0. More relevant works of system (1.2) can refer to ([16,17,18]). Furthermore, many scholars consider the situation which the growth or death of cells is influenced by external factors, that is, the logistic sources contains nonlocal growth term. Negreanu and Tello [19] proposed the following model
{ut=Δu−χ∇⋅(um∇v)+u(a0−a1u−a2∫Ωudx),x∈Ω,t>0,−Δv+λv=f+u,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, | (1.2) |
where Ω⊂Rn is a smooth bounded domain, if m=1 and a1>2χ+|a2|, it is shown that the solution of (1.2) satisfies limt→∞‖u−a0a1+a2‖L∞(Ω)=0. when the second is replaced by vt=Δv−v+uγ and the logistic source term is uα(1−σ|Ω|∫Ωuβdx) with α≥1,β>1 and m=γ=σ|Ω|=1, Bian [20] proved that the system (1.2) possesses a unique global strong solution which is uniformly bounded in whole space under either γ+1≤σ<1+2βn or σ<γ+1<2(σ+β)n+2+nn+2. In reference [21], when γ+m≤α<1+2βn or n+42−β<α<γ+m, the system (1.2) admits a uniformly bounded global classical solution. For the logistic source term is replaced by u(a0−a1uα+a2∫Ωuαdx), Negreanu [22] obtained the global-in-time existence of classical solutions and the convergence to steady state (a1α0(a1−a2|Ω|)−1α,(a1α0(a1−a2|Ω|)−1α)γ) under assumptions that α,γ≥1,m>1,α+1>m+γ,a1>0 and a1−a2|Ω|>0. Moreover, when the logistic source term is replaced by uσ(a0−a1u−a2∫Ωuβdx), Ren [23] proved that system (1.2) possesses a unique global classical solution in three different cases, namely parabolic-elliptic, fully parabolic, and parabolic-parabolic-elliptic. For more chemotaxis systems with nonlocal terms, we can find the literature works ([24,25,26,27]). The chemotactic signal is produced directly by cells in the classical Keller–Segel system, yet the signal generation undergoes intermediate stages in some realistic biological processes [28]. The related models can be described as follows:
{ut=∇⋅(D(u)∇u)−∇⋅(S(u)∇v)+μ(u−uγ),x∈Ω,t>0,τvt=Δv−v+w,x∈Ω,t>0,τwt=Δw−w+u,x∈Ω,t>0,∂νu=∂νv=∂νw=0,x∈∂Ω,t>0,u(x,0)=u0(x),τv(x,0)=τv0(x),τw(x,0)=τw0(x),x∈Ω, | (1.3) |
where u,v,w represent the density of cells, the density of chemical substances and the concentration of indirect signal, respectively. For τ=0,D(s)≥a0(s+1)α,|S(s)|≤b0(1+s)β−1 for all s≥0 with a0,b0>0,α,β∈R, Li in [29] have obtained the nonnegative classical solution (u,v,w) is global in time and bounded for β≤γ−1. Moreover, if μ satisfies some suitable conditions, the solution (u,v,w) converges to (1,1,1) in L∞-norm as t→∞. When the signal generation is in a nonlinear form, it also has been shown that the boundedness and large time behaviors of classical solutions in [30]. For τ=1,D(u)=1,S(u)=u, Zhang in [28] proved that if γ>n4+12, the solution is globally bounded; if μ>0 is sufficiently large, (u,v,w) satisfies ‖u(⋅,t)−1‖L∞(Ω)+‖v(⋅,t)−1‖L∞(Ω)+‖w(⋅,t)−1‖L∞(Ω)→0 as t→∞. For D(u)≥D1uα,S(u)=u with D1>0,α∈R,γ≥2 and μ>0, Wu in [31] proved the global existence and boundedness of solutions if the assumption αn+γ2>12 holds with n≥3. When D,S∈C2([0,∞)) satisfying D(s)≥a0(s+1)−α,0≤S(s)≤b0(s+1)β for a0,b0>0 and logistic source term is replaced by b−μsγ for all s,b≥0,γ≥1, Wang [32] obtained the global boundedness of solutions in four cases: the self-diffusion dominates the cross-diffusion; the logistic source suppresses the cross-diffusion for μ>0 sufficiently large; the logistic dampening balances the cross-diffusion; the self-diffusion and the logistic source both balance the cross-diffusion with μ>0 suitably large. If D,S are smooth functions satisfying D(s)≥a0(s+1)α,|S(s)|≤b0s(s+1)β−1 for all s>0 and the logistic source term is μ(s−sγ) with s≥0,γ>1, Zhang in [33] showed that the system (1.3) possesses a globally bounded classical solution (u,v,w).
The above systems only discuss the forms of direct or indirect generation of chemical signals, but do not discuss systems with nonlocal source terms and indirect signal production.Thereafter, inspired by reference [22], considering the growth or death of cells is influenced by external factors, this article added the production of chemical signals goes through intermediate stages and studied the following fully parabolic chemotaxis system with a nonlocal growth term and indirect signal production
{ut=∇⋅(D(u)∇u)−∇⋅(u∇v)+f(u),x∈Ω,t>0,vt=Δv+w−v,x∈Ω,t>0,wt=Δw+u−w,x∈Ω,t>0,∂u∂n=∂v∂n=∂w∂n=0,x∈∂Ω,t>0,u(x,0)=u0(x),v(x,0)=v0(x),w(x,0)=w0(x),x∈Ω, | (1.4) |
in a bounded domain Ω⊂Rn(n≥3) with smooth boundary ∂Ω. The diffusion function D(u) satisfies
D∈C2([0,∞)) and D(u)≥D1uγ,u>0, | (1.5) |
where D1 is a positive constant and γ∈R. And f(u) is the logistic function, which satisfies
f(u)=u(a0−a1uσ+a2∫Ωuσdx) | (1.6) |
with
σ≥max{1,−γ} and a0∈R,ai(i=1,2)>0,a1−a2|Ω|>0, | (1.7) |
the initial data fulfill
{u0∈C0(ˉΩ),u0≥0,u0≢0,v0∈W1,∞(ˉΩ),v0≥0,w0∈W1,∞(ˉΩ),w0≥0. | (1.8) |
Under these assumptions, our main results on the global boundedness and large time behavior of solutions to system (1.4) are as follows:
Theorem 1. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. Suppose that functions D(u),f(u) and parameters σ,ai(i=0,1,2) satisfy (1.5)-(1.7) with 4n+γσ>3−σ1+σ. For any nonnegative initial data (u0,v0,w0) evolve from (1.8), the system (1.4) possesses a global classical solution
(u,v,w)∈(C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax)))3, | (1.9) |
which is bounded in Ω×(0,∞).
Remark 1. When a2=0, the result in Theorem 1 is consistent with that in [31] with S(u)=u in system (1.3), which needs μ>0. If a2>0, system (1.4) does not require any restrictions on a0. These findings suggest that the nonlocal term a2∫Ωuσdx could play a crucial role in ensuring the global existence and boundedness of solutions in (1.4).
Remark 2. This work extends the study in [22] to the system with nonlinear diffusion and indirect signal production. Our results show that the nonlinear diffusion mechanism and nonlocal logistic sources have an inhibitory effect on the blow-up solution.
Theorem 2. Assume the conditions in Theorem 1 hold. Let (u,v,w) be the solution of (1.4) obtained in Theorem 1, then there exist some κ>0 and C=C(a0,a1,a2,σ,γ,|Ω|)>0 such that
‖u(⋅,t)−u∗‖L∞(Ω)+‖v(⋅,t)−v∗‖L∞(Ω)+‖w(⋅,t)−w∗‖L∞(Ω)≤Ce−κt |
for all t>0, where κ=14(n+2),u∗=v∗=w∗=(a0a1−a2|Ω|)1σ.
The framework structure of this article is as follows. In Sect.2, we give the local existence of the solution in (1.4) and show several related inequalities. In Sect.3, we consider the global boundedness of solutions for problem (1.4) under some suitable conditions and prove Theorem 1. In Sect.4, we give a lower bound for u via comparison [34]. In Sect.5, we obtain the asymptotic behavior of (1.4) by constructing Lyapunov functions, thus completing the proof of Theorem 2.
In this section, we present several important lemmas that will be used in the following sections. First, we will state the local-in-time existence result of solutions for problem (1.4), which can be proved by adapting well-established approaches for parabolic-parabolic chemotaxis models (see [35]).
Lemma 1. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary, D(u),f(u), and initial data (u0,v0,w0) satisfy (1.5), (1.6), and (1.8), respectively. Then there exist Tmax∈(0,∞] and a tripe (u,v,w) of nonnegative functions
(u,v,w)∈(C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax)))3 |
which solves (1.4) in the classical sense. Moreover, if Tmax<∞, we can see that
limt↗Tmaxsup‖u(⋅,t)‖L∞(Ω)=∞. | (2.1) |
We next state a lemma which guarantees L1-boundedness of u,v and w.
Lemma 2. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and the parameters σ,ai(i=0,1,2) satisfy (1.5)-(1.7). Assume that (u,v,w) is the solution of (1.4). Then there exist constants Mi(i=1,2,3,4) such that
‖u(⋅,t)‖L1(Ω)≤M1:=max{∫Ωu0dx,(a0a1−a2|Ω|)1σ|Ω|}, | (2.2) |
‖v(⋅,t)‖L1(Ω)≤M2:=max{∫Ωv0,M3}, | (2.3) |
‖w(⋅,t)‖L1(Ω)≤M3:=max{∫Ωw0,M1}, | (2.4) |
for all t∈(0,Tmax). Moreover,
∫t+τt∫Ωuσ+1dxdt≤M4=(a1−a2|Ω|)−1M1(1+a0τ), | (2.5) |
for all τ∈(0,Tmax) and t∈(0,Tmax−τ).
Proof. Integrating the first equation of the system (1.4) over Ω, we have
ddt∫Ωudx=∫Ωu(a0−a1uσ+a2∫Ωuσdx)dx=a0∫Ωudx−a1∫Ωuσ+1dx+a2∫Ωu(∫Ωuσdx)dx. | (2.6) |
By applying Hölder's inequality, we derive
∫Ωu(∫Ωuσdx)dx≤(∫Ωuσ+1dx)1σ+1|Ω|σσ+1(∫Ωuσ⋅σ+1σdx)σσ+1|Ω|1σ+1=|Ω|∫Ωuσ+1dx, | (2.7) |
therefore
a2∫Ωu(∫Ωuσdx)dx≤a2|Ω|∫Ωuσ+1dx. |
Assuming (1.7) holds, applying the Hölder's inequality again yields
ddt∫Ωudx≤a0∫Ωudx−(a1−a2|Ω|)∫Ωuσ+1dx≤a0∫Ωudx−a1−a2|Ω||Ω|σ(∫Ωudx)σ+1, | (2.8) |
for all t∈(0,Tmax).
On an ordinary differential equation(ODE) comparison, this implies that
∫Ωudx≤M1, | (2.9) |
where M1:=max{∫Ωu0dx,(a0a1−a2|Ω|)1σ|Ω|}. Integrating the third equation in (1.4) over Ω yields
ddt∫Ωwdx=∫Ωudx−∫Ωwdx, for all t∈(0,Tmax), | (2.10) |
then
ddt∫Ωwdx+∫Ωwdx=∫Ωudx≤M1, for all t∈(0,Tmax). | (2.11) |
So (2.2) and (2.11) imply (2.4). By the same method, using the second equation of the system (1.4) and (2.4), we obtain (2.3). Moreover, integrating (2.8) upon (t,t+τ) for t∈(0,Tmax−τ), and using (2.9), we have
∫t+τt∫Ωuσ+1dxdt≤M4, for all t∈(0,Tmax−τ), |
where M4=(a1−a2|Ω|)−1M1(1+a0τ).
We provide the well-known Neumann heat semigroup theory without proof for our subsequent work (see Ref. [36,37]).
Lemma 3. Assume that λ∈{0,1},p∈[1,∞],q∈(1,∞) and θ∈(0,1). Then there exists positive constant c1 such that for all u∈D(Aθ),A:=−Δ+λ,
‖u‖Wm,p(Ω)≤c1‖(−Δ+1)θu‖Lq(Ω), | (2.12) |
if
m−np≤2θ−nq. |
If in addition q≥p, then there exist some constants c2>0 and α>0 such that for all u∈Lp(Ω),
‖Aθe−tAu‖Lq(Ω)≤c2t−θ−n2(1p−1q)e−αt‖u‖Lp(Ω), | (2.13) |
where the associated diffusion semigroup {e−tA}t≥0 maps Lp(Ω) into D(Aθ). Moreover, for any p∈(1,∞) and ε>0, there exist c3>0 and μ>0 such that
‖AθetΔ∇⋅u‖Lp(Ω)≤c3t−θ−12−εe−μt‖u‖Lp(Ω) | (2.14) |
is valid for all Rn-Valued u∈Lp(Ω).
We recall the Gagliardo–Nirenberg interpolation inequality (see Ref. [38] for detail), which will be used frequently in the proof of our main results.
Lemma 4. Let u∈W1,q(Ω)∩Lp(Ω) and h≥1,p∈(0,h), where Ω⊂Rn(n≥3) is a smooth bounded domain. Then there exists a constant c4>0 such that
‖u‖Lh(Ω)≤c4(‖∇u‖λLq(Ω)‖u‖1−λLp(Ω)+‖u‖Lp(Ω)), | (2.15) |
where
1h=λ(1q−1n)+(1−λ)1p, |
and λ∈(0,1) satisfies
λ=np−nh1−nq+np. |
Finally, we give the following lemma from [39], which is also important for our proof.
Lemma 5. Let T>0,c5>0,c6>0 and τ∈(0,T). Assume that the function z:[0,T)→[0,∞) is absolutely continuous and such that the following inequality holds:
z′(t)+c5z(t)≤c(t), | (2.16) |
for a.e t∈(0,T), where c∈L1loc([0,T)) is a nonnegative function satisfying
∫t+τtc(s)ds≤c6, | (2.17) |
for all t∈[0,T−τ).
Then
z(t)≤max{z(0)+c6,c6c5τ+2c6}, | (2.18) |
for a.e t∈(0,T).
In this section, we will give the proof of global existence and boundedness of solutions to system (1.4). Some necessary estimations are needed. We first give an inequality involving ∫Ωwp+1dx and then use this inequality to establish the Lp-estimate for w.
Lemma 6. Assume that (u,v,w) is the solution of (1.4) on [0,Tmax) as in Lemma 2.1 and p>1. Then we have the following inequality
ddt∫Ωwp+1dx+p∫Ωwp+1dx≤pp∫Ωup+1dx, | (3.1) |
for all t∈(0,Tmax).
Proof. Multiplying the third equation of (1.4) by wp, integrating over Ω and with the help of Young's inequality, we obtain
1p+1ddt∫Ωwp+1dx=∫Ωwp(Δw+u−w)dx=−p∫Ωwp−1|∇w|2dx+∫Ωwpudx−∫Ωwp+1dx≤1p+1∫Ωwp+1dx+ppp+1∫Ωup+1dx−∫Ωwp+1dx=−pp+1∫Ωwp+1dx+ppp+1∫Ωup+1dx. |
In addition, in order to obtain the proof of Theorem 1, a key step is to derive the upper bounds of ‖w(⋅,t)‖Lσ+1(Ω) and ‖∇v(⋅,t)‖Lβ(Ω). We will expand the proof from the following lemmas.
Lemma 7. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). Assume that (u,v,w) is the solution of (1.4). Then for each β∈(1,n(σ+1)n−(σ+1)), there exist positive constants M5 and M6 such that
‖w(⋅,t)‖Lσ+1(Ω)≤M5,‖∇v(⋅,t)‖Lβ(Ω)≤M6,for allt∈(0,Tmax). | (3.2) |
Proof. Using Lemma 6 for p=σ, we have
1σ+1ddt∫Ωwσ+1dx≤−σσ+1∫wσ+1dx+σσσ+1∫Ωuσ+1dx. | (3.3) |
Taking z(t):=∫Ωwσ+1dx,t∈(0,Tmax), and c(t):=σσ∫Ωuσ+1(⋅,t)dx,t∈(0,Tmax), we obtain
z′(t)+σz(t)≤c(t), for all t∈(0,Tmax). | (3.4) |
According to (2.5), one implies
∫t+τtc(t)dt=σσ∫t+τt∫Ωuσ+1dxdt≤C1, for all t∈(0,Tmax−τ), | (3.5) |
where C1=σσM4>0.
In view of (3.5) and Lemma 5, this yields that
∫Ωwσ+1dx≤max{∫Ωwσ+10+C1,C1στ+2C1}, |
thus, we have
‖w(⋅,t)‖Lσ+1(Ω)≤M5, | (3.6) |
where M5=max{∫Ωwσ+10+C1,C1στ+2C1}1σ+1.
Next, we obtain the Lβ-bound of ∇v by applying semigroup arguments (see, for example, [36,37]). First, using the variation of constants formula for the second equation of system (1.4) indicates
v(⋅,t)=et(Δ−1)v0+∫t0e(t−s)(Δ−1)w(⋅,s)ds. | (3.7) |
Choosing τ<1 and setting A:=−Δ+1,λ=1,m=1,q=σ+1,β∈(1,n(σ+1)n−(σ+1)) in Lemma 3, which makes θ∈(12+n2(σ+1)−n2β,1). Then there exist positive constants ε,γ such that
‖v(⋅,t)‖W1,β(Ω)≤C2‖Aθv(⋅,t)‖Lσ+1(Ω)≤C2‖Aθe−tAv0‖Lσ+1(Ω)+C2∫t0‖Aθe−(t−s)A‖w(⋅,s)‖Lσ+1(Ω)ds≤C2t−θ−n2(1σ+1−1σ+1)e−εt‖v0‖Lσ+1(Ω)+C2∫t0(t−s)−θe−γ(t−s)‖w(⋅,s)‖Lσ+1(Ω)ds≤C2t−θ‖v0‖Lσ+1(Ω)+C2M5∫t0(t−s)−θe−γ(t−s)ds, | (3.8) |
for all t∈(τ,Tmax). So
‖v(⋅,t)‖W1,β(Ω)≤C2τ−θ‖v0‖Lσ+1(Ω)+C2Γ(1−θ):=C3, | (3.9) |
where C2,C3 represent different positive constants, and when (1−θ)>0,Γ(1−θ)>0. Therefore, we obtain that ‖∇v(⋅,t)‖Lβ(Ω) is bounded for all t∈(0,Tmax).
Next we will calculate the first equation of system (1.4) to obtain an inequality for ∫Ωupdx by applying the classical Green's formula, Young's inequality, and Hölder's inequality.
Lemma 8. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). Assume that (u,v,w) is the solution of (1.4). Then for any p>max{2,γ},σ>−γ there exist C4,C5>0 depend on p, such that the following inequality
ddt∫Ωupdx+∫Ωupdx≤−(a1−a2|Ω|)∫Ωup+σdx+C4∫Ω|∇v|2(p+σ)σ+γdx+C5 | (3.10) |
holds, for all t∈(0,Tmax).
Proof. Multiplying the first equation in (1.4) by up−1 and integrating over Ω by parts, we have
1pddt∫Ωupdx=∫Ωup−1∇⋅(D(u)∇u)dx−∫Ωup−1∇⋅(u∇v)dx+∫Ωup(a0−a1uσ+a2∫Ωuσdx)dx≤−(p−1)∫Ωup−2D(u)|∇u|2dx+(p−1)∫Ωup−1∇u⋅∇vdx+a0∫Ωupdx−a1∫Ωup+σdx+a2∫Ωupdx∫Ωuσdx≤−(p−1)∫ΩD1up+γ−2|∇u|2dx+(p−1)∫Ωup−1∇u⋅∇vdx+a0∫Ωupdx−a1∫Ωup+σdx+a2∫Ωupdx∫Ωuσdx, | (3.11) |
for all t∈(0,Tmax).
Using Young's inequality to the second term on the right side of (3.11) yields
(p−1)∫Ωup−1∇u⋅∇vdx≤(p−1)D12∫Ωup+γ−2|∇u|2dx+(p−1)2D1∫Ωup−γ|∇v|2dx≤(p−1)D12∫Ωup+γ−2|∇u|2dx+C4p∫Ω|∇v|2(p+σ)σ+γdx+a1−a2|Ω|2p∫Ωup+σdx,for allt∈(0,Tmax). | (3.12) |
According to Young's inequality, the combination of (3.11) and (3.12) leads to
1pddt∫Ωupdx+1p∫Ωupdx≤−(p−1)D12∫Ωup+γ−2|∇u|2dx+(a0+1p)∫Ωupdx+C4p∫Ω|∇v|2(p+σ)σ+γdx+(a1−a2|Ω|2p−a1)∫Ωup+σdx+a2∫Ωupdx∫Ωuσdx≤−(p−1)D12∫Ωup+γ−2|∇u|2dx+(a0+1p)∫Ωupdx+C4p∫Ω|∇v|2(p+σ)σ+γdx+(a1−a2|Ω|2p−a1)∫Ωup+σdx+a2|Ω|∫Ωup+σdx≤−(p−1)D12∫Ωup+γ−2|∇u|2dx+C4p∫Ω|∇v|2(p+σ)σ+γdx−(a1−a2|Ω|p)∫Ωup+σdx+C5p, | (3.13) |
where we note that our assumption a1−a2|Ω|>0 and p>2 warrant that
a1−a2|Ω|−3(a1−a2|Ω|)2p>0. |
Therefore, the following inequality
(a0+1p)∫Ωupdx≤(a1−a2|Ω|−3(a1−a2|Ω|)2p)∫Ωup+σdx+C5p |
holds by Young's inequality. Then (3.10) is obtained.
In order to obtain the Lp boundedness of u, we need to estimate the second term on the right side of (3.10), thus an inequality for ∫Ω|∇v|2qdx has to be required.
Lemma 9. Let (u,v,w) be a solution of (1.4), then for any q>2 there exist positive constants C6 and C7 such that
ddt∫Ω|∇v|2qdx+2q∫Ω|∇v|2qdx+2q−2q∫Ω|∇|∇v|q|2dx≤(p−1)∫Ωwp+1dx+C6∫Ω|∇v|(2q−2)(p+1)p−1dx+C7,for allt∈(0,Tmax). | (3.14) |
Proof. According to the second equation of (1.4), we can obtain
12qddt∫Ω|∇v|2qdx=∫Ω|∇v|2q−2∇v⋅∇vtdx=∫Ω|∇v|2q−2∇v⋅∇(Δv+w−v)dx=∫Ω|∇v|2q−2∇v⋅∇wdx−∫Ω|∇v|2qdx+∫Ω|∇v|2q−2∇v⋅∇(Δv)dx, | (3.15) |
for all t∈(0,Tmax).
Now we estimate the first and third terms of (3.15), where an important inequality (3.10) in reference [10] as follows is needed
12∫∂Ω|∇v|2q−2∂|∇v|2∂ndx≤(q−1)2q2∫Ω|∇|∇v|q|2dx+C8, | (3.16) |
with some C8>0.
Using (3.16), we can infer that
∫Ω|∇v|2q−2∇v⋅∇(Δv)dx=12∫Ω|∇v|2q−2Δ|∇v|2dx−∫Ω|∇v|2q−2|D2v|2dx=−12∫Ω∇|∇v|2q−2⋅∇|∇v|2dx+12∫∂Ω|∇v|2q−2⋅∂|∇v|2∂ndx−∫Ω|∇v|2q−2|D2v|2dx=−q−12∫Ω|∇v|2q−4⋅|∇|∇v|2|2dx+12∫∂Ω|∇v|2q−2⋅∂|∇v|2∂ndx−∫Ω|∇v|2q−2|D2v|2dx≤−2(q−1)q2|∇|∇v|q|2dx+q−12q2∫Ω|∇|∇v|q|2dx−∫Ω|∇v|2q−2|D2v|2dx+C8≤−3(q−1)2q2|∇|∇v|q|2dx−∫Ω|∇v|2q−2|D2v|2dx+C8, | (3.17) |
for all t∈(0,Tmax), where we have used Δ|∇v|2=2∇v⋅∇(Δv)+2|D2v|2. Applying Young's inequality and |Δv|2≤n|D2v|2, we get
∫Ω|∇v|2q−2∇v⋅∇wdx=−(q−1)∫Ωw|∇v|2(q−2)∇|∇v|2⋅∇vdx−∫Ωw|∇v|2q−2Δvdx≤q−1q2∫Ω|∇|∇v|q|2dx+(q−1)∫Ωw2|∇v|2q−2dx+√n∫Ωw|∇v|2q−2|D2v|dx≤q−1q2∫Ω|∇|∇v|q|2dx+(q−1)∫Ωw2|∇v|2q−2dx+∫Ω|∇v|2q−2|D2v|2dx+n4∫Ωw2|∇v|2q−2dx≤q−1q2∫Ω|∇|∇v|q|2dx+((q−1)+n4)∫Ωw2|∇v|2q−2dx+∫Ω|∇v|2q−2|D2v|2dx, | (3.18) |
for all t∈(0,Tmax).
And we also have
((q−1)+n4)∫Ωw2|∇v|2q−2dx≤p−12q∫Ωwp+1dx+C62q∫Ω|∇v|(2q−2)(p+1)p−1dx, | (3.19) |
for all t∈(0,Tmax). Substituting (3.17), (3.18), and (3.19) into (3.15), we have
12qddt∫Ω|∇v|2qdx≤−q−1q2∫Ω|∇|∇v|q|2dx−∫Ω∇v|2qdx+p−12q∫Ωwp+1dx+C62q∫Ω|∇v|(2q−2)(p+1)p−1dx+C8, | (3.20) |
for all t∈(0,Tmax). Then we readily derive (3.14), where C7:=2qC8.
Combining Lemma 6 and Lemma 8 with Lemma 9 and using Young's inequality, the following inequality will be gained. And next we will establish the boundedness of ‖u(⋅,t)‖Lp(Ω), ‖∇v(⋅,t)‖L2p(Ω) and ‖w(⋅,t)‖Lp+1(Ω).
Lemma 10. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and the parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). Assume that (u,v,w) is the solution of (1.4). Then for any p>max{2,γ},q>2, there exist positive constants M7,M8 and M9 such that
‖u(⋅,t)‖Lp(Ω)≤M7,‖∇v(⋅,t)‖L2q(Ω)≤M8,‖w(⋅,t)‖Lp+1(Ω)≤M9, | (3.21) |
for all t∈(0,Tmax).
Proof. If σ>1 or σ≥−γ, we combine (3.1), (3.10) with (3.14) and apply Young's inequality to obtain
ddt(∫Ωupdx+∫Ω|∇v|2qdx+∫Ωwp+1dx)+∫Ωupdx+2q∫Ω|∇v|2qdx+∫Ωwp+1dx+2q−2q∫Ω|∇|∇v|q|2dx≤pp∫Ωup+1dx−(a1−a2|Ω|)∫Ωup+σdx+C4∫Ω|∇v|2(p+σ)σ+γdx+C6∫Ω|∇v|(2q−2)(p+1)p−1dx+C5+C7≤C4∫Ω|∇v|2(p+σ)σ+γdx+C6∫Ω|∇v|(2q−2)(p+1)p−1dx+C9, for all t∈(0,Tmax). | (3.22) |
If σ=1, we have the similar inequality and omit it here. In order to obtain the boundedness above, we need to estimate the right two terms of (3.22). By employing reference [38] and Gagliardo–Nirenberg inequality, we deduce that
C4∫Ω|∇v|2(p+σ)σ+γdx=C4‖|∇v|q‖2(p+σ)q(σ+γ)L2(p+σ)q(σ+γ)(Ω)=C4‖|∇v|q‖α1qLα1q(Ω)≤C10(‖∇|∇v|q‖λ1L2(Ω)‖|∇v|q‖1−λ1Lβq(Ω)+‖|∇v|q‖Lβq(Ω))α1q=C10(‖∇|∇v|q‖λ1L2(Ω)‖∇v‖q(1−λ1)Lβ(Ω)+‖∇v‖qLβ(Ω))α1q≤C10(‖∇|∇v|q‖λ1L2(Ω)Mq(1−λ1)6+Mq6)α1q≤C10max{Mq(1−λ1)6,Mq6}α1q⋅(‖∇|∇v|q‖λ1L2(Ω)+1)α1q≤C10max{Mq(1−λ1)6,Mq6}α1q⋅2α1q−1(‖∇|∇v|q‖α1λ1qL2(Ω)+1)≤˜C10(‖∇|∇v|q‖h1L2(Ω)+1)≤q−1q‖∇|∇v|q‖2L2(Ω)+C11, for all t∈(0,Tmax), | (3.23) |
and
C6∫Ω|∇v|(2q−2)(p+1)p−1dx=C6‖|∇v|q‖(2q−2)(p+1)q(p−1)L(2q−2)(p+1)q(p−1)(Ω)=C6‖|∇v|q‖α2qLα2q(Ω)≤C12(‖∇|∇v|q‖λ2L2(Ω)‖|∇v|q‖1−λ2Lβq(Ω)+‖|∇v|q‖Lβq(Ω))α2q=C12(‖∇|∇v|q‖λ2L2(Ω)‖∇v‖q(1−λ2)Lβ(Ω)+‖∇v‖qLβ(Ω))α2q≤˜C12(‖∇|∇v|q‖h2L2(Ω)+1)≤q−1q‖∇|∇v|q‖2L2(Ω)+C13, for all t∈(0,Tmax), | (3.24) |
where
α1=2(p+σ)σ+γ, α2=(2q−2)(p+1)p−1,λ1=qβ−qα1qβ−(12−1n), λ2=qβ−qα2qβ−(12−1n),h1=α1β−1qβ−(12−1n), h2=α2β−1qβ−(12−1n). | (3.25) |
And for large p>1, it will be shown that there exists constant q>1 such that λ1,λ2∈(0,1) and h1,h2<2 hold in (3.25).
To ensure that λ1,λ2∈(0,1) and h1,h2<2, we can choose suitable parameters such that
α1,α2>β and q>α12−βn,q>α22−βn. | (3.26) |
According to continuity argument, we discuss the case β∈(1,n(σ+1)n−(σ+1)), which is inserted into (3.26) to have
p>n(σ+1)2(n−(σ+1))(σ+γ)−σ,q>n(σ+1)2(n−(σ+1))+1, | (3.27) |
and
p+σσ+γ−σ+1n−(σ+1)<q<p+12+(σ+1)(p−1)2(n−(σ+1)). | (3.28) |
Thus if
p+σσ+γ−σ+1n−(σ+1)<p+12+(σ+1)(p−1)2(n−(σ+1)), | (3.29) |
holds, then (3.28) is true. And 4n+γσ>3−σ1+σ implies (3.29). Then for any p>p0, we can choose suitable q such that (3.27) and (3.28) are fulfilled, where p0:=max{2,γ,n(σ+1)2(n−(σ+1))(σ+γ)−σ}.
Hence, a combination of (3.22)-(3.24) entails that
ddt(∫Ωup+∫Ω|∇v|2q+∫Ωwp+1)dx+∫Ωupdx+2q∫Ω|∇v|2qdx+∫Ωwp+1dx≤C14, | (3.30) |
for all t∈(0,Tmax). By employing the Grönwall's inequality for (3.30), the desired result (3.21) is gained.
Lemma 11. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). Assume (u,v,w) is the solution of (1.4), then there exists a positive constant M10 such that
‖∇v(⋅,t)‖L∞(Ω)≤M10,for allt∈(0,Tmax). | (3.31) |
Proof. (3.31) can be deduced by applying the semigroup arguments (see [36,37] for details) to the third and second equations of (1.4).
Next, by using Lemma 11 and the standard Alikakos–Moser iteration, we establish the L∞ bound of u.
Lemma 12. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). Assume that (u,v,w) is the solution of (1.4). Then for any p>max{2,γ},σ>−γ, there exists a positive constant M11>0 such that
‖u(⋅,t)‖L∞(Ω)≤M11,for allt∈(0,Tmax). | (3.32) |
Proof. For any p>max{2,γ}, multiplying the first equation of (1.4) by up−1, integrating over Ω, and using Hölder's inequality, one obtains
1pddt∫Ωupdx=∫Ωup−1∇⋅(D(u)∇u)dx−∫Ωup−1∇⋅(u∇v)dx+∫Ωup−1f(u)dx≤−(p−1)∫Ωup−2D(u)|∇u|2dx+(p−1)∫Ωup−1∇u⋅∇vdx+a0∫Ωupdx−a1∫Ωup+σdx+a2∫Ωup∫Ωuσdxdx≤−(p−1)D1∫Ωup+γ−2|∇u|2dx+(p−1)∫Ωup−1∇u⋅∇vdx+a0∫Ωupdx−(a1−a2|Ω|)∫Ωup+σdx | (3.33) |
Once more employing Young's inequality, we have
ddt∫Ωupdx+∫Ωupdx≤C15. | (3.34) |
Integrating (3.34) yields that
∫Ωupdx≤C16,C16:=max{∫Ωup0dx,C15}. | (3.35) |
Then, we can prove the following inequality by using Alikakos–Moser iteration (see [16,40] for details)
‖u(⋅,t)‖L∞(Ω)≤M11, for all t∈(0,Tmax). | (3.36) |
Now, we complete the proof of Theorem 1.
Proof of Theorem 1. Along with Lemma 1 part 2, this proves that Tmax=∞ and the standard parabolic regularity makes sure that (u,v,w) is bounded for (x,t)∈Ω×(0,∞). Hence the desired result of Theorem 1 is obtained.
To further prove the asymptotic behavior of the solution, we are going to estimate the lower bound for u and provide some lemmas as follows.
Lemma 13. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u) and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7). If (u,v,w) is the solution of (1.4) and(u0,v0,w0) satisfy (1.8), then there exists a constant positive C17 such that
limt→∞sup‖Δv(⋅,t)‖L∞(Ω)≤M12. | (4.1) |
Proof. We split the proof into two steps:
Step 1. Fix θ0∈(0,1) and set m=2,p=∞ in Lemma 3, we verify limt→∞sup‖Aθ0w(⋅,t)‖Lp(Ω)≤C17 for any p>max{2,γ,n(σ+1)2(n−(σ+1))(σ+γ)−σ} with some C17=C17(θ0,p,λ,γ,σ)>0.
Review of (3.21), we know that there exists suitably large t0>0 satisfying
‖u(⋅,t)‖Lp(Ω)≤M7, for all t≥t0. | (4.2) |
In view of the variation-of-constants representation, we get
w(⋅,t)=e(t−t0)(Δ−1)w(⋅,t0)+∫tt0e(t−s)(Δ−1)u(⋅,s)ds=e−(1−λ)(t−t0)e−(t−t0)Aw(⋅,t0)+∫tt0e−(1−λ)(t−s)e−(t−s)Au(⋅,s)ds |
for all t≥t0. Recalling that A=−Δ+λ in reference [36] and applying the (4.2), then there exist constants C18,C19, we can derive
‖Aθ0w(⋅,t)‖Lp(Ω)≤e−(1−λ)(t−t0)‖Aθ0e−(t−t0)Aw(⋅,t0)‖Lp(Ω)+∫tt0e−(1−λ)(t−s)‖Aθ0e−(t−s)Au(⋅,s)‖Lp(Ω)ds≤C18e−(1−λ)(t−t0)(t−t0)−θ0‖w(⋅,t0)‖Lp(Ω)+C19∫tt0e−(1−λ)(t−s)(t−s)−θ0‖u(⋅,s)‖Lp(Ω)ds≤C18e−(1−λ)(t−t0)(t−t0)−θ0‖w(⋅,t0)‖Lp(Ω)+C20 |
for all t≥t0, where C20:=M7C19∫∞0e−(1−λ)ττ−θ0dτ<∞ with θ0∈(0,1). Thereafter
limt→∞sup‖Aθ0w(⋅,t)‖L∞(Ω)≤C21, |
and there exists t0>0 large enough fulfilling
‖Aθ0w(⋅,t)‖L∞(Ω)≤2C21, for all t≥t0. | (4.3) |
Step 2. The solution v satisfies (4.1).
Let us fix any θ1∈(1,2) on the condition that θ1−1<θ0<1 and choose p>max{2,γ,nn−2(σ+γ)−σ} satisfying p>n2(θ1−1), thus
2θ1−np>2θ1−2(θ1−1)=2. | (4.4) |
Then following the same procedure as Step 1 and invoking (2.11), (4.3), and reference [36], we estimate
‖v(⋅,t)‖W2,∞≤C22‖Aθ1v(⋅,t)‖Lp(Ω)≤C22e−(1−λ)(t−t0)‖Aθ1e−(t−t0)Av(⋅,t0)‖Lp(Ω)+C22∫tt0e−(1−λ)(t−s)‖Aθ1e−(t−s)Aw(⋅,s)‖Lp(Ω)ds≤C23e−(1−λ)(t−t0)(t−t0)−θ1‖v(⋅,t0)‖Lp(Ω)+C24 |
for all t≥t0, where C24:=2C21C′22∫∞0e−(1−λ)ττ−(θ1−θ0)dτ<∞ with θ1−1<θ0<1. This implies (4.1) for some M12>0. In accordance with (4.1), we can pick t0>0 suitably large satisfying
‖Δv(⋅,t)‖L∞(Ω)≤2M12, for all t≥t0. | (4.5) |
With the help of the comparison principle [34], a positive lower bound for u is hereafter addressed after some suitable waiting time.
Lemma 14. Let Ω⊂Rn(n≥3) be a bounded domain with smooth boundary. The functions D(u),f(u), and parameters σ,ai(i=1,2) satisfy (1.5)-(1.7) with the positive number p being as defined in (4.4). Assume that (u,v,w) is the solution of (1.4) and the initial data fulfills (1.8). Then there exists a constant M13 such that
limt→∞inf(infx∈Ωu(x,t))≥(1−a0−a1−2M12a1)1σ:=M13, | (4.6) |
Proof. Dealing with the first equation in (1.4), a combination of (3.34), and (4.5) yields
ut=∇⋅(D(u)∇u)−∇⋅(u∇v)+u(a0−a1uσ+a2∫Ωuσdx)=∇⋅(D(u)∇u)−∇u⋅∇v−uΔv+a0u−a1uσ+1+a2u∫Ωuσdx≥∇⋅(D(u)∇u)−∇u⋅∇v−2M12u+a0u−a1uσ+1+a2u∫Ωuσdx≥∇⋅(D(u)∇u)−∇u⋅∇v+(a0−2M12)u−a1uσ+1, |
for all x∈Ω and t≥t0. Moreover, let the function y(t)∈C1([t0,∞)) be defined by
{y′(t)=(a0−2M12)⋅y(t)−a1yσ+1(t),t≥t0,y(t0)=infx∈Ωu(x,t0). |
Applying the comparison argument, we gain
u(x,t)≥y(t),y(t)=C25eσ(a0−2M12)t+C26, for all x∈Ω and t≥t0, |
and
y(t)→(1−a0−a1−2M12a1)1σ, as t→∞. |
Which implies that
lim inft→∞(infx∈Ωu(x,t))≥lim inft→∞infy(t)≥(1−a0−a1−2M12a1)1σ. |
The key to proving Theorem 2 relies on seeking so-called Lyapunov functions. Thus in this section, we will construct the appropriate Lyapunov functions in the following lemmas to obtain the large-time behavior of the solution (u∗,v∗,w∗) of the system (1.4).
Lemma 15. (Lemma 3.1 in [41]) Let f:(1,∞)→[0,∞) be uniformly continuous such that ∫∞1f(t)dt<∞, then
f(t)→0,ast→∞. | (5.1) |
Lemma 16. Let functions D(u) and f(u) and parameters σ and ai(i=1,2) satisfy (1.5)-(1.7). Then for any classical solution (u,v,w) of (1.4) in Ω×(t0,∞) verifying
supt∈[t0,∞)(‖u(⋅,t)‖L∞(Ω)+‖v(⋅,t)‖W1,∞(Ω))≤M14, |
with some t0≥0 and M14>0.
Then there exist ˉθ∈(0,1) and C27=C27(M14,D1,σ,γ,|Ω|,‖∇u(⋅,t0+2)‖C(ˉΩ))>0 such that
‖∇u(⋅,t)‖C(ˉΩ)≤C27((t−t0)ˉθ+1),t>t0+2. |
Proof. First, using the Neumann heat semigroup estimate [34] to the third and the second equations in system (1.4) we can derive that ‖∇v‖L∞(Ω) and ‖∇w‖L∞(Ω) is bounded in t∈(0,Tmax). According to Theorem 1, the system (1.4) possesses a global bounded classical solution. Thus, we can find t∈[t0,∞) such that
supt∈[t0,∞)(‖u(⋅,t)‖L∞(Ω)+‖v(⋅,t)‖W1,∞(Ω))≤M14. |
And because of (u,v,w)∈(C0(ˉΩ×[0,Tmax))∩C2,1(ˉΩ×(0,Tmax)))3, applying a proof method similar to proposition 3.4 in Reference [15], we can obtain that there exists ˉθ∈(0,1) such that ‖∇u(⋅,t)‖C(ˉΩ)≤C27((t−t0)ˉθ+1) holds.
Lemma 17. Assume functions D(u) and f(u) and parameters σ and ai(i=1,2) satisfy (1.5)-(1.7) and (u,v,w) is the solution of (1.4). The initial data (u0,v0,w0) evolves from (1.8). Then there exists a positive constant H0∈[M13,M11] such that functions E and F are defined by
E(t):=∫Ω(u−u∗−u∗lnuu∗)dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗2∫Ω(w−w∗)2dx, | (5.2) |
F(t):=∫Ω(u−u∗)2dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−w∗)2dx, | (5.3) |
where t>0,u∗=v∗=w∗=(a1a1−a2|Ω|)1σ, and we have
ddtE(t)≤−F(t),for allt>0. | (5.4) |
Proof. Note that
E(t)=∫Ω(u−u∗−u∗lnuu∗)dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗2∫Ω(w−w∗)2dx:=E1(t)+E2(t)+E3(t). | (5.5) |
By system (1.4) and the assumption in Theorem 2, applying Young's inequality, we deduce that
E′1(t)=∫Ωu−u∗uutdx=∫Ωu−u∗u(∇⋅(D(u)∇u)−∇⋅(u∇v)+u(a0−a1uσ+a2∫Ωuσdx))dx=−u∗∫Ω|∇u|2u2D(u)dx+u∗∫Ω∇u⋅∇vudx+∫Ω(u−u∗)(a0−a1uσ+a2∫Ωuσdx)dx≤−u∗2∫ΩD1uγ|∇u|2u2dx+u∗2∫Ω1D1uγ|∇v|2dx−a1∫Ω(u−u∗)(uσ−uσ∗)dx+a2∫Ω(u−u∗)dx∫Ω(uσ−uσ∗)dx≤H0u∗2∫Ω|∇v|2dx−a1∫Ω(u−u∗)(uσ−uσ∗)dx+a2∫Ω(u−u∗)dx∫Ω(uσ−uσ∗)dx, | (5.6) |
where a0=a1uσ∗−a2∫Ωuσ∗dx, H0 is obtained by u with a lower bound.
A simple calculation yields
(u−u∗)(uσ−uσ∗)≥uσ−1∗(u−u∗)2, |
and hence
−a1∫Ω(u−u∗)(uσ−uσ∗)dx≤−a1uσ−1∗∫Ω(u−u∗)2dx. | (5.7) |
We treat the last integral in (5.6) by estimating the integrated function in a pointwise way. Suppose (1.5) is valid, according to reference [15], and the Mean value theorem ensure
a2∫Ω(u−u∗)dx∫Ω(uσ−uσ∗)dx≤a2(∫Ω(u−u∗)2dx)12|Ω|12(∫Ω(uσ−uσ∗)2dx)12|Ω|12≤a2|Ω|∫Ω(u−u∗)2dx+a2|Ω|∫Ω(uσ−uσ∗)2dx≤a2|Ω|∫Ω(u−u∗)2dx+a2|Ω|∫Ω(σ(u+u∗)σ−1|u−u∗|)2dx≤a2|Ω|∫Ω(u−u∗)2dx+a2|Ω|σ2(M1+lu∗)2σ−2∫Ω(u−u∗)2dx, | (5.8) |
where l>0, together with (5.7)-(5.9), we can conclude that
E′1(t)≤H0u∗2∫Ω|∇v|2dx−a1uσ−1∗∫Ω(u−u∗)2dx+a2|Ω|∫Ω(u−u∗)2dx+a2|Ω|σ2(M1+lu∗)2σ−2∫Ω(u−u∗)2dx. | (5.9) |
Then we use Young's equality to estimate E2(t) and E3(t),
E′2(t)=H0u∗2∫Ωvt(v−v∗)dx=H0u∗2∫Ω(v−v∗)Δvdx+H0u∗2∫Ωw(v−v∗)dx−H0u∗2∫Ωv(v−v∗)dx=−H0u∗2∫Ω|∇v|2dx+H0u∗2∫Ω(v−v∗)(w−w∗)dx−H0u∗2∫Ω(v−v∗)2dx≤−H0u∗2∫Ω|∇v|2dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−v∗)2dx−H0u∗2∫Ω(v−v∗)2dx=−H0u∗2∫Ω|∇v|2dx−H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−v∗)2dx=−H0u∗2∫Ω|∇v|2dx−H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−w∗)2dx | (5.10) |
and
E′3(t)=H0u∗∫Ωwt(w−w∗)dx=H0u∗∫Ω(w−w∗)Δwdx+H0u∗∫Ωu(w−w∗)dx−H0u∗∫Ωw(w−w∗)dx≤−H0u∗∫Ω|∇w|2dx−H0u∗2∫Ω(w−w∗)2dx+H0u∗2∫Ω(u−w∗)2dx=−H0u∗∫Ω|∇w|2dx−H0u∗2∫Ω(w−w∗)2dx+H0u∗2∫Ω(u−u∗)2dx. | (5.11) |
Combing E′1(t), E′2(t) and E′3(t), we obtain
ddtE(t)+a1uσ−1∗∫Ω(u−u∗)2dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−w∗)2dx≤(a2|Ω|+a2|Ω|σ2(M1+lu∗)2σ−2+H0u∗2)∫Ω(u−u∗)dx. | (5.12) |
Since
a2|Ω|+a2|Ω|σ2(M1+lu∗)2σ−2+H0u∗2a1uσ−1∗≤a2|Ω|+a2|Ω|σ2(M1+lu∗)2σ−2+H0u∗2a1uσ−1∗−a2|Ω|uσ−1∗=a2|Ω|+a2|Ω|σ2(M1+l(a0a1−a2|Ω|)1σ)2σ−2+H02(a0a1−a2|Ω|)1σa0⋅(a1−a2|Ω|a0)1σ→0 as a1→∞, |
there is ~a1>0 such that
a1uσ−1∗>a2|Ω|+a2|Ω|σ2(M1+lu∗)2σ−2+H0u∗2+1 |
provided a1>~a1. This in conjunction with (5.13) entails that
ddtE(t)+∫Ω(u−u∗)2dx+H0u∗4∫Ω(v−v∗)2dx+H0u∗4∫Ω(w−w∗)2dx≤0, |
therefore
ddtE(t)≤−∫Ω(u−u∗)2dx−H0u∗4∫Ω(v−v∗)2dx−H0u∗4∫Ω(w−w∗)2dx≤−F(t), | (5.13) |
where a1−a2|Ω|>0. Lemma 5.3 is proved.
We are now in a positive to prove our main result.
Proof of Theorem 2. Integrating (5.13) from t0 to ∞, we have
∫∞t0∫Ω(u−u∗)2dxdt+∫∞t0∫Ω(v−v∗)2dxdt+∫∞t0∫Ω(w−w∗)2dxdt≤∞, |
According to the standard parabolic regularity theory[42], with the global boundedness of (u,v,w), we can see that there exists ϑ∈(0,1) and C28>0 such that
‖u‖C2+ϑ,1+ϑ2(ˉΩ×[t,t+1])+‖v‖C2+ϑ,1+ϑ2(ˉΩ×[t,t+1])+‖w‖C2+ϑ,1+ϑ2(ˉΩ×[t,t+1])≤C28, | (5.14) |
for all t>1. This clearly implies that ∫Ω(u(⋅,t)−u∗)2dx+∫Ω(v(⋅,t)−v∗)2dx+∫Ω(w(⋅,t)−w∗)2dx is uniformly continuous with respect to t≥t0 provided t0>1. Therefore, we infer from Lemma 15 that the following inequality
∫Ω(u(⋅,t)−u∗)2dx+∫Ω(v(⋅,t)−v∗)2dx+∫Ω(w(⋅,t)−w∗)2dx→0 as t→∞. | (5.15) |
holds.
The Gagliardo–Nirenberg inequality says
‖ϖ‖L∞(Ω)≤CGN‖ϖ‖nn+2W1,∞(Ω)‖ϖ‖2n+2L2(Ω) for all ϖ∈W1,∞(Ω). | (5.16) |
Using ϖ as u−u∗,v−v∗ and w−w∗ in (5.16), respectively, we have from (5.14) and (5.15) that
‖u(⋅,t)−u∗‖L∞(Ω)+‖v(⋅,t)−v∗‖L∞(Ω)+‖w(⋅,t)−w∗‖L∞(Ω)→0, as t→∞. | (5.17) |
We next defined φ:(0,∞)→R and
φ(s):=s−u∗−u∗lnsu∗,s>0. |
By the Taylor expansion with s∈(0,∞), there is ξ(s)∈(0,1) such that
φ(s):=φ(u∗)+φ′(u∗)(s−u∗)+φ″(ξs+(1−ξ)u∗)2(s−u∗)2=u∗2(ξs+(1−ξ)u∗)2(s−u∗)2, s>0. |
Obviously, φ(s)≥0, and use L'Hôpital's rule to see
lims→u∗φ(s)(s−u∗)2=lims→u∗u∗2(ξs+(1−ξ)u∗)2=12u∗. |
Furthermore, we can pick t1>0 such that
14u∗(u−u∗)2≤φ(u(⋅,t))≤1u∗(u−u∗)2, for all t≥t1, | (5.18) |
namely,
14u∗∫Ω(u(⋅,t)−u∗)2dx≤E1(t)≤1u∗∫Ω(u(⋅,t)−u∗)2dx, for all t≥t1, | (5.19) |
In view of (5.13) and (5.19), we have
ddtE(t)≤−F(t)≤−12E(t), for all t≥t1, |
thus, we get by the Gronwall inequality that
E(t)≤E(t1)e−12(t−t1), for all t≥t1, |
Together with (5.18), we obtain the estimate
‖u(⋅,t)−u∗‖2L2(Ω)+‖v(⋅,t)−v∗‖2L2(Ω)+‖w(⋅,t)−w∗‖2L2(Ω)≤C29e−12(t−t1), | (5.20) |
for all t≥t1, with C29=C29(a0,a1,a2,σ,γ,|Ω|)>0. By using the Gagliardo–Nirenberg inequality with Lemma 16 and (5.14)-(5.15),
‖u(⋅,t)−u∗‖L∞(Ω)≤C30‖∇u(⋅,t)‖nn+2L∞(Ω)‖u(⋅,t)−u∗‖2n+2L2(Ω)≤C31((t−t1)ˉθ+1)e−(t−t1)2(n+2)≤C32e−(t−t1)4(n+2), t>t1+2 | (5.21) |
with C30=C30(|Ω|),C31=C31(a0,a1,a2,σ,γ,|Ω|,‖∇u(⋅,t1)‖L∞(Ω))>0,C32=C32(C31,ˉθ). Similarly, an application of the Gagliardo–Nirenberg inequality with combining (3.31) and (5.20) indicates that
‖v(⋅,t)−v∗‖L∞(Ω)≤C33e−(t−t1)4(n+2), t>t1+2 | (5.22) |
‖w(⋅,t)−w∗‖L∞(Ω)≤C34e−(t−t1)4(n+2), t>t1+2 | (5.23) |
for some C33=C33(a0,a1,a2,σ,γ,|Ω|,‖∇v(⋅,t1)‖L∞(Ω))>0, and C34=C34(a0,a1,a2,σ,γ,|Ω|,‖∇w(⋅,t1)‖L∞(Ω))>0.
A combination of (5.21)-(5.23), gives us
‖u(⋅,t)−u∗‖L∞(Ω)+‖v(⋅,t)−v∗‖L∞(Ω)+‖w(⋅,t)−w∗‖L∞(Ω)≤Ce−(t−t1)4(n+2), |
with t≥t1 and C=C(a0,a1,a2,σ,γ,|Ω|)>0, which completes the proof of Theorem 2.
In this paper, we considered that the growth or death of cells is influenced by external factors and the production of chemical signals goes through intermediate stages, and thus we investigate a fully parabolic chemotaxis system with a nonlocal growth term and indirect signal production. The work is carried out under the condition of spatial dimension n≥3, when the initial data, the diffusion function, the logistic source term and related parameters satisfy certain conditions, the global boundedness of solutions to system (1.4) is proved by applying the maximum principle, variation-of constants formula, Neumann heat semigroup estimation, Young's inequality, Gagliardo–Nirenberg inequality and so on. In addition, by constructing appropriate Lyapunov functions, we obtained the asymptotic behavior of solutions.
Min Jiang, Dandan Liu and Rengang Huang: Methodology; Min Jiang and Dandan Liu: Writing-original draft; Dandan Liu and Rengang Huang: Writing-review and editing.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors are indebted to the editor and anonymous reviewers for their kind assistance in providing insightful comments, suggestions, and valuable references. Additionally, the authors extend appreciation to reviewers for thoroughly reviewing the manuscript, engaging in fruitful discussions, and bringing to attention certain errors throughout the course of this study. The research of the first author was supported by the Project of Guizhou Minzu University under Grant No.16yjrcxm002 and the Science Research Foundation Project of Guizhou Minzu University (GZMUZK[2023]YB12).
The authors declare there is no conflicts of interest.
[1] | E. F. Keller, L. A. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol., 26 (1970), 399–415. |
[2] |
K. Osaki, T. Tsujikawa, A. Yagi, M. Mimura, Exponential attractor for a chemotaxis-growth system of equations, Nonlinear Anal., 51 (2002), 119–144. https://doi.org/10.1016/s0362-546x(01)00815-x doi: 10.1016/s0362-546x(01)00815-x
![]() |
[3] |
J. I. Tello, M. Winkler, A chemotaxis system with logistic source, Commun. Partial Differ. Equ., 32 (2007), 849–877. https://doi.org/10.1080/03605300701319003 doi: 10.1080/03605300701319003
![]() |
[4] |
M. Winkler, Chemotaxis with logistic source: very weak global solutions and their boundedness properties, J. Math. Anal. Appl., 348 (2008), 708–729. https://doi.org/10.1016/j.jmaa.2008.07.071 doi: 10.1016/j.jmaa.2008.07.071
![]() |
[5] |
X. Cao, Boundedness in a quasilinear parabolic-parabolic Keller-Segel system with logistic source, J. Math. Anal. Appl., 412(2014), 181-188. https://doi.org/10.1016/j.jmaa.2013.10.061 doi: 10.1016/j.jmaa.2013.10.061
![]() |
[6] | K. Osaki, A. Yagi, Finite dimensional attractor for one-dimensional Keller-Segel equations, Funkcial. Ekvac., 44 (2001), 441–469. |
[7] | M. A. Herrero, J. J. Velázquez, A blow-up mechanism for a chemotaxis model, Ann. Sc. Norm. Super. Pisa Cl. Sci., 24 (1997), 633–683. |
[8] | T. Nagai, T. Senba, K. Yoshida, Application of the Trudinger-Moser inequality to a parabolic system of chemotaxis, Funkcial. Ekvac., 40 (1997), 411–433. |
[9] |
M. Winkler, Finite-time blow-up in the higher-dimensional parabolic-parabolic Keller-Segel system, J. Math. Pure. Appl., 100 (2013), 748–767. https://doi.org/10.1016/j.matpur.2013.01.020 doi: 10.1016/j.matpur.2013.01.020
![]() |
[10] |
S. Ishida, K. Seki, T. Yokota, Boundedness in quasilinear Keller-Segel systems of parabolic-parabolic type on non-convex bounded domains, J. Differ. Equ., 256 (2014), 2993–3010. https://doi.org/10.1016/j.jde.2014.01.028 doi: 10.1016/j.jde.2014.01.028
![]() |
[11] |
X. Cao, Large time behavior in the logistic Keller-Segel model via maximal Sobolev regularity, Discrete Contin. Dyn. Syst. Ser. B., 22 (2017), 3369–3378. https://doi.org/10.3934/dcdsb.2017141 doi: 10.3934/dcdsb.2017141
![]() |
[12] |
Q. Wang, J. Yang, F. Yu, Boundedness in logistic keller-segel models with nonlinear diffusion and sensitivity functions, Discrete Contin. Dyn. Syst., 37 (2017), 5021-5036. https://doi.org/10.3934/dcds.2017216 doi: 10.3934/dcds.2017216
![]() |
[13] |
J. Zheng, Boundedness of solutions to a quasilinear parabolic-parabolic Keller-Segel system with a logistic source, J. Math. Anal. Appl., 431 (2015), 867–888. https://doi.org/10.1016/j.jmaa.2015.05.071 doi: 10.1016/j.jmaa.2015.05.071
![]() |
[14] |
M. Zhuang, W. Wang, S. Zheng, Boundedness in a fully parabolic chemotaxis system with logistic-type source and nonlinear production, Nonlinear Anal-Real., 47 (2019), 473–483. https://doi.org/10.1016/j.nonrwa.2018.12.001 doi: 10.1016/j.nonrwa.2018.12.001
![]() |
[15] |
M. Ding, W. Wang, S. Zhou, S. Zheng, Asymptotic stability in a fully parabolic quasilinear chemotaxis model with general logistic source and signal production, J. Differ. Equ., 268 (2020), 6729–6777. https://doi.org/10.1016/j.jde.2019.11.052 doi: 10.1016/j.jde.2019.11.052
![]() |
[16] |
Y. Tao, M. Winkler, Boundedness in a quasilinear parabolic-parabolic Keller-Segel system with subcritical sensitivity, J. Differ. Equ., 252 (2012), 692–715. https://doi.org/10.1016/j.jde.2011.08.019 doi: 10.1016/j.jde.2011.08.019
![]() |
[17] |
D. Liu, Y. Tao, Boundedness in a chemotaxis system with nonlinear signal production, Appl. Math. J. Chin. Univ. Ser. B., 31 (2016), 379–388. https://doi.org/10.1007/s11766-016-3386-z doi: 10.1007/s11766-016-3386-z
![]() |
[18] |
M. Winkler, A critical blow-up exponent in a chemotaxis system with nonlinear signal production, Nonlinearity, 31 (2018), 2031–2056. https://doi.org/10.1088/1361-6544/aaaa0e doi: 10.1088/1361-6544/aaaa0e
![]() |
[19] |
M. Negreanu, J. I. Tello, On a competitive system under chemotactic effects with non-local terms, Nonlinearity, 26 (2013), 1083–1103. https://doi.org/10.1088/0951-7715/26/4/1083 doi: 10.1088/0951-7715/26/4/1083
![]() |
[20] |
S. Bian, L. Chen, E. A. Latos, Chemotaxis model with nonlocal nonlinear reaction in the whole space, Discrete Contin. Dyn. Syst., 38 (2018), 5067–5083. https://doi.org/10.3934/dcds.2018222 doi: 10.3934/dcds.2018222
![]() |
[21] | E. A. Latos, Nonlocal reaction preventing blow-up in the supercritical case of chemotaxis, arXiv preprint arXiv: 2011.10764, 2020. https://doi.org/10.48550/arXiv.2011.10764 |
[22] |
M. Negreanu, J. I. Tello, A. M. Vargas, On a fully parabolic chemotaxis system with nonlocal growth term, Nonlinear Anal., 213 (2021), 112518. https://doi.org/10.1016/j.na.2021.112518 doi: 10.1016/j.na.2021.112518
![]() |
[23] |
G. Ren, Global boundedness and asymptotic behavior in an attraction-repulsion chemotaxis system with nonlocal terms, Z. Angew. Math. Phys., 73 (2022), 200. https://doi.org/10.1007/s00033-022-01832-7 doi: 10.1007/s00033-022-01832-7
![]() |
[24] |
T. B. Issa, R. B. Salako, Asymptotic dynamics in a two-species chemotaxis model with non-local terms, Discrete Contin. Dyn. Syst. Ser. B., 22 (2017), 3839–3874. https://doi.org/10.3934/dcdsb.2017193 doi: 10.3934/dcdsb.2017193
![]() |
[25] |
T. B. Issa, W. Shen, Persistence, coexistence and extinction in two species chemotaxis models on bounded heterogeneous environments, J. Dyn. Differ. Equ., 31 (2019), 1839–1871. https://doi.org/10.1007/s10884-018-9686-7 doi: 10.1007/s10884-018-9686-7
![]() |
[26] |
P. Zheng, On a parabolic-elliptic Keller-Segel system with nonlinear signal production and nonlocal growth term, Dynam. Part. Differ. Eq., 21 (2024), 61–76. https://doi.org/10.4310/DPDE.2024.v21.n1.a3 doi: 10.4310/DPDE.2024.v21.n1.a3
![]() |
[27] |
Y. Chiyo, F. G. Düzgün, S. Frassu, G. Viglialoro, Boundedness through nonlocal dampening effects in a fully parabolic chemotaxis model with sub and superquadratic growth, Appl. Math. Opt., 89 (2024), 9. https://doi.org/10.1007/s00245-023-10077-3 doi: 10.1007/s00245-023-10077-3
![]() |
[28] |
W. Zhang, P. Niu, S. Liu, Large time behavior in a chemotaxis model with logistic growth and indirect signal production, Nonlinear Anal. Real Word Appl., 50 (2019), 484–497. https://doi.org/10.1016/j.nonrwa.2019.05.002 doi: 10.1016/j.nonrwa.2019.05.002
![]() |
[29] |
D. Li, Z. Li, Asymptotic behavior of a quasilinear parabolic-elliptic-elliptic chemotaxis system with logistic source, Z. Angew. Math. Phys., 73 (2022), 1–17. https://doi.org/10.1007/s00033-021-01655-y doi: 10.1007/s00033-021-01655-y
![]() |
[30] |
C. Wang, Y. Zhu, X. Zhu, Long time behavior of the solution to a chemotaxis system with nonlinear indirect signal production and logistic source, Electron. J. Qual. Theo., 2023 (2023), 1–21. https://doi.org/10.14232/ejqtde.2023.1.11 doi: 10.14232/ejqtde.2023.1.11
![]() |
[31] |
S. Wu, Boundedness in a quasilinear chemotaxis model with logistic growth and indirect signal production, Acta. Appl. Math., 176 (2021), 1–14. https://doi.org/10.1007/s10440-021-00454-x doi: 10.1007/s10440-021-00454-x
![]() |
[32] |
W. Wang, A quasilinear fully parabolic chemotaxis system with indirect signal production and logistic source, J. Math. Anal. Appl., 477 (2019), 488–522. https://doi.org/10.1016/j.jmaa.2019.04.043 doi: 10.1016/j.jmaa.2019.04.043
![]() |
[33] |
W. Zhang, S. Liu, P. Niu, Asymptotic behavior in a quasilinear chemotaxis-growth system with indirect signal production, J. Math. Anal. Appl., 486 (2020), 123855. https://doi.org/10.1016/j.jmaa.2020.123855 doi: 10.1016/j.jmaa.2020.123855
![]() |
[34] |
M. Winkler, Global asymptotic stability of constant equilibria in a fully parabolic chemotaxis system with strong logistic dampening, J. Differ. Equ., 257 (2014), 1056–1077. https://doi.org/10.1016/j.jde.2014.04.023 doi: 10.1016/j.jde.2014.04.023
![]() |
[35] |
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
![]() |
[36] |
D. Horstmann, M. Winkler, Boundedness vs. blow-up in a chemotaxis system, J. Differ. Equ., 215 (2005), 52–107. https://doi.org/10.1016/j.jde.2004.10.022 doi: 10.1016/j.jde.2004.10.022
![]() |
[37] |
M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model, J. Differ. Equ., 248 (2010), 2889–2905. https://doi.org/10.1016/j.jde.2010.02.008 doi: 10.1016/j.jde.2010.02.008
![]() |
[38] | L. Nirenberg, On elliptic partial differential equations, Ann. Scuola. Norm-Sci., 13 (1959), 115–162. |
[39] |
C. Stinner, C. Surulescu, M. Winkler, Global weak solutions in a PDE-ODE system modeling multiscale cancer cell invasion, SIAM. J. Comput., 46 (2014), 1969–2007. https://doi.org/10.1137/13094058X doi: 10.1137/13094058X
![]() |
[40] |
C. Mu, L. Wang, P. Zheng, Q. Zhang, Global existence and boundedness of classical solutions to a parabolic-parabolic chemotaxis system, Nonlinear Anal-Real., 14 (2013), 1634–1642. https://doi.org/10.1016/j.nonrwa.2012.10.022 doi: 10.1016/j.nonrwa.2012.10.022
![]() |
[41] |
X. Bai, M. Winkler, Equilibration in a fully parabolic two-species chemotaxis system with competitive kinetics, Indiana. U. Math. J., 65 (2016), 553–583. https://doi.org/10.1512/iumj.2016.65.5776 doi: 10.1512/iumj.2016.65.5776
![]() |
[42] | O. A. Ladyzhenskaia, V. A. Solonnikov, N. N. Ural'tseva, Linear and quasi-linear equations of parabolic type, AMS, 1968. |