In this paper, we consider the incompressible chemotaxis-Navier-Stokes equations with logistic source in spatial dimension two. We first show a blow-up criterion and then establish the global existence of classical solutions to the system for the Cauchy problem under some rough conditions on the initial data.
Citation: Yina Lin, Qian Zhang, Meng Zhou. Global existence of classical solutions for the 2D chemotaxis-fluid system with logistic source[J]. AIMS Mathematics, 2022, 7(4): 7212-7233. doi: 10.3934/math.2022403
[1] | Zihang Cai, Chao Jiang, Yuzhu Lei, Zuhan Liu . Global classical solution of the fractional Nernst-Planck-Poisson-Navier- Stokes system in R3. AIMS Mathematics, 2024, 9(7): 17359-17385. doi: 10.3934/math.2024844 |
[2] | Jianxia He, Ming Li . Existence of global solution to 3D density-dependent incompressible Navier-Stokes equations. AIMS Mathematics, 2024, 9(3): 7728-7750. doi: 10.3934/math.2024375 |
[3] | Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363 |
[4] | Xin Zhao, Xin Liu, Jian Li . Convergence analysis and error estimate of finite element method of a nonlinear fluid-structure interaction problem. AIMS Mathematics, 2020, 5(5): 5240-5260. doi: 10.3934/math.2020337 |
[5] | Xindan Zhou, Zhongping Li . Global bounded solution of a 3D chemotaxis-Stokes system with slow p-Laplacian diffusion and logistic source. AIMS Mathematics, 2024, 9(6): 16168-16186. doi: 10.3934/math.2024782 |
[6] | Özlem Kaytmaz . The problem of determining source term in a kinetic equation in an unbounded domain. AIMS Mathematics, 2024, 9(4): 9184-9194. doi: 10.3934/math.2024447 |
[7] | Harald Garcke, Kei Fong Lam . Global weak solutions and asymptotic limits of a Cahn–Hilliard–Darcy system modelling tumour growth. AIMS Mathematics, 2016, 1(3): 318-360. doi: 10.3934/Math.2016.3.318 |
[8] | Qasim Khan, Anthony Suen, Hassan Khan, Poom Kumam . Comparative analysis of fractional dynamical systems with various operators. AIMS Mathematics, 2023, 8(6): 13943-13983. doi: 10.3934/math.2023714 |
[9] | Ning Duan, Xiaopeng Zhao . On global well-posedness to 3D Navier-Stokes-Landau-Lifshitz equations. AIMS Mathematics, 2020, 5(6): 6457-6463. doi: 10.3934/math.2020416 |
[10] | Jiali Yu, Yadong Shang, Huafei Di . Existence and nonexistence of global solutions to the Cauchy problem of thenonlinear hyperbolic equation with damping term. AIMS Mathematics, 2018, 3(2): 322-342. doi: 10.3934/Math.2018.2.322 |
In this paper, we consider the incompressible chemotaxis-Navier-Stokes equations with logistic source in spatial dimension two. We first show a blow-up criterion and then establish the global existence of classical solutions to the system for the Cauchy problem under some rough conditions on the initial data.
The coupled chemotaxis-Navier-Stokes model with logistic source terms is as following: QT=(0,T]×R2
{nt+u⋅∇n=DnΔn−∇⋅(nχ(c)∇c)+κn−μn2,(x,t)∈QT,ct+u⋅∇c=DcΔc−g(c)n,(x,t)∈QT,ut+(u⋅∇)u−∇P=DuΔu+n∇Φ,(x,t)∈QT,∇⋅u=0,(x,t)∈QT. | (1.1) |
Here the unknows are the concentration of bacteria n=n(x,t):QT→R+; the oxygen concentration c=c(x,t):QT→R+; the given vector field u=u(x,t):QT→R2; and the pressure of the fluid P=P(x,t):QT→R. We denote the corresponding diffusion coefficients for the cells, the oxygen and the fluid by Dn,Dc and Du.
Apart from that, χ(c) represents the chemotactic sensitivity, and g(c) is the consumption rate of oxygen. In [19], it is shown that the functions g(c) and χ(c) are constants at large c and rapidly approach zero below some critical c∗. Moreover, the experimentalists in [19] used multiples of the Heaviside step function to model χ(⋅) and g(⋅). The scalar valued function Φ is also given.
The model (1.1) was proposed in [12], which in order to describe the movement of bacteria in response to the presence of a chemical signal substance (oxygen or another nutrient) in their fluid environment. κ is the effective growth rate of the population and μ controlling death by overcrowding. For more the logistic term related, see [10,11,24,35].
In recent years, the study of the dynamics of solutions to (1.1) has attracted many researchers. Finite time blow-up phenomena is one of the most important dynamical problems in (1.1). One can refer to [3,30]. It is worth pointing that the logistic term κn−μn2 of (1.1) can avoid blow-up phenomena in [9].
To make the system (1.1) be well-posed, we add some initial conditions as follows:
(n,c,u)|t=0=(n0(x),c0(x),u0(x)),x∈R2. | (1.2) |
Besides, we also have the corresponding chemotaxis-only subsystem of (1.1) on letting u≡0, that is, the system
{nt=DnΔn−∇⋅(nχ(c)∇c)+κn−μn2,ct=DcΔc−c+n, | (1.3) |
is a variant of the classical Keller-Segel model. Up to now, one notices that the homogeneity of model (1.3) may be undermined by the cross-diffusive term −∇⋅(nχ(c)∇c) and even enforce blow-up of solutions [1]. Although we are facing great challenges, there are still numerous analytic results which mainly fixed attention on the local and global solvability of corresponding initial(-boundary)-value problems in either bounded or unbounded domains Ω, under diverse technical conditions on χ(⋅) and g(⋅).
As for initial boundary value problem (1.3) in higher space dimensions, Winkler [22] proved the global existence of a small solution to (1.3) and that it asymptotically behaves like the solution of a decoupled system of linear parabolic equations. On the other hand, a result concerning that blow-up behavior occurring for some radially symmetric positive initial data in higher dimension was recently obtained in [23]. And one can see a relevant reference in [25].
If system (1.3) is coupled with Navier-Stokes equations, we have the following model
{nt+u⋅∇n=DnΔn−∇⋅(nχ(c)∇c)+κn−μn2,ct+u⋅∇c=DcΔc−c+n,ut+u⋅∇u=DuΔu−∇P−n∇ϕ+f(x,t). | (1.4) |
As far as we know, results on smooth global solvability could be established for the two-dimensional version of (1.4) whenever μ is positive ([18]), while a similar statement could be derived when μ≥23 in three dimension at least for a Stokes simplification of (1.4) in which the nonlinear convective term u⋅∇u is neglected [17]. In a recent paper [31], Winkler reveals that whenever ω>0, requiring that
κmin{μ,μ32+ω}<η |
with some η=η(ω)>0, and that f satisfies a suitable assumption on ultimate smallness, is sufficient to ensure that each of these generalized solutions becomes eventually smooth and classical.
For system (1.1), when κ=μ=0, [6] demonstrates that there exists a global weak solutions for the 2D chemotaxis-Stokes equations, i.e., the nonlinear convective term u⋅∇u is removed in the fluid equation of (1.1), by making use of quasi-energy functionals associated with (1.1), under the following conditions on χ(⋅) and g(⋅):
χ(c)>0,χ′(c)≥0,g(0)=0,g′(c)>0,d2dc2(g(c)χ(c))<0, | (1.5) |
χ(c)≥0,χ′(c)≥0,g(c)≥0,g′(c)≥0, |
d2dc2(g(c)χ(c))≤0,χ′(c)g(c)+χ(c)g′(c)χ(c)>0, | (1.6) |
for the fully chemotaxis-Navier-Stokes system (1.1) without any smallness conditions on the initial data.
Furthermore, for the smooth initial data, Chae et al. establish the global existence of smooth solutions [3] by assuming that χ(c) and g(c) satisfy
χ(c)≥0,χ′(c)≥0,g(c)≥0,g′(c)≥0,g(0)=0, | (1.7) |
and that there exists a constant η such that
supc≥0|χ(c)−ηg(c)|<ϵfor a sufficiently smallϵ>0, | (1.8) |
which is taken away in [4]. For the full 3D chemotaxis-Navier-Stokes system, the global classical solution near constant steady states and the global weak solutions under the special situation that χ(⋅) precisely coincides with a fixed multiple of g(⋅) are constructed in [6] and [3], respectively.
In the case of κ=μ=0, Winkler establishes the existence of global classical solutions in 2D bounded convex domains Ω in [26]. Then he [27] asserts that a solution of the two-dimensional chemotaxis-Navier-Stokes system stabilizes to the spatially uniform equilibrium (ˉn,0,0) with ¯n0=1|Ω|∫Ωn0(x)dx in the sense of L∞(Ω). If κ≥0, μ>0, in a bounded smooth domain Ω⊂R3, J. Lankeit [12] shows that after some waiting time weak solutions constructed become smooth and finally converge to the semi-trivial steady state (κμ,0,0). More related paper on the bounded domain case, for the 2D case, one may refer to [16,20,21] and for the 3D case, one also refer to [2,16,21,28,29]. Additionally, under some strong structural assumptions on χ and g, the global existence of weak solutions is proved as well as their eventual smoothness and stabilization to the 3D version of the chemotaxis-Navier-Stokes system is established in [28] and [29].
Recently, Duan et al. [7] established the global existence of weak solutions and classical solutions for both the Cauchy problem and the initial-boundary value problem supplemented with some initial data. In addition, Wu et al. [32] improved the results by taking more careful calculations than those of [7].
The main purpose of this paper is to demonstrate the global existence of classical solutions to (1.1) for the Cauchy problem. We first introduce the corresponding blowup criterion.
Now, it is position to state our main theorems in this article.
Theorem 1.1. Let m≥3. Assume that χ(⋅), g(⋅)∈Cm(R) with g(0)=0, and that ‖∇lΦ‖L∞(R2)<∞ for 1≤|l|≤m.
(A1) Then there exists T∗>0, the maximal time of existence, such that, if the initial data (n0,c0,u0)∈Hm−1(R2)×Hm(R2)×Hm(R2;R2), then there exists a unique classical solution (n,c,u) to system (1.1)-(1.2) satisfying for any T<T∗
(n,c,u)∈L∞(0,T;Hm−1(R2)×Hm(R2)×Hm(R2;R2)) |
and
(∇n,∇c,∇u)∈L2(0,T;Hm−1(R2)×Hm(R2)×Hm(R2;R2)). |
(A2) Moreover, if the maximal time of existence T∗<∞, then
∫T∗0‖∇c(τ)‖2L∞(R2)dτ=∞ | (1.10) |
The proof of A1 in Theorem 1.1 is standard, one can refer to [3,33]. We will focus on proving part A2.
Theorem 1.2. Let m≥3. Under the assumptions of Theorem 1.1, then system (1.1)-(1.2) admits a unique global-in-time classical solution (n,c,u) satisfying for any T>0
(n,c,u)∈L∞(0,T;Hm−1(R2)×Hm(R2)×Hm(R2;R2)) |
and
(∇n,∇c,∇u)∈L2(0,T;Hm−1(R2)×Hm(R2)×Hm(R2;R2)). |
Before we are going to prove the main results, we give an important proposition, which can be found in [14].
Let
X0={n0∈L1∩L2, n0>0,∇√c0∈L2, c0∈L1∩L∞, c0>0,u0∈L2}. |
Proposition 1.1. Let the triple (n0,c0,u0)∈X0, ∇Φ∈L∞(R2) and χ(⋅), g(⋅)∈Cm(R) with m≥3 and g(0)=0. Then, system (1.1) has a unique global solution (n,c,u) such that
n∈L∞(R+;L1(R2))∩L∞loc(R+;L2(R2))∩L2loc(R+;H1(R2));c∈L∞(R+;L1(R2)∩L∞(R2))∩L∞loc(R+;H1(R2))∩L2loc(R+;H2(R2));u∈L∞(R+;L1(R2))∩L2loc(R+;H1(R2)). |
The rest of this article is organized as follows. Let's briefly establish a blow-up criterion in Section 2, discuss the Cauchy problem in Section 3.
Now, we show the proof of the blow-up criterion for the fluid chemotaxis equations.
Proof. Above all, we take the L2 estimate of n into account. Multiplying n on both sides of (1.1)1 and integrating in spaces, by the fact that χ is continuous and c is uniformly bounded, we deduce
12ddt‖n‖2L2+Dn‖∇n‖2L2+μ‖n‖3L3≤C‖χ(c)n∇c‖L2‖∇n‖L2+κ‖n‖2L2≤C‖∇c‖2L∞‖n‖2L2+14‖∇n‖2L2+κ‖n‖2L2. |
Next, testing −Δc to both sides of (1.1)2 and making use of the integration by parts, it follows that
12ddt‖∇c‖2L2+Dc‖Δc‖2L2≤‖∇(u⋅∇c)−(u⋅∇)∇c‖L2‖∇c‖L2+C‖g(c)n‖L2‖Δc‖L2≤C‖∇u‖L∞‖∇c‖2L2+C‖n‖2L2+14‖Δc‖2L2. |
Similary, taking the L2 inner product −Δu to both sides of (1.1)3 and using the integration by parts, we also deduce
12ddt‖∇u‖2L2+Du‖Δu‖2L2≤‖∇((u⋅∇)u)−(u⋅∇)∇u‖L2‖∇u‖L2+C‖n‖L2‖Δu‖L2≤C‖∇u‖L∞‖∇u‖2L2+C‖n‖2L2+C‖Δu‖2L2. |
From all the estimates, it means that
ddt(‖n‖2L2+‖∇c‖2L2+‖∇u‖2L2)+(‖∇n‖2L2+‖Δc‖2L2+‖Δu‖2L2)≤C(1+‖∇u‖L∞+‖∇c‖2L∞)(‖n‖2L2+‖∇c‖2L2+‖∇u‖2L2). |
An application of Gronwall's inequality yields that
sup(‖n‖2L2+‖∇c‖2L2+‖∇u‖2L2)+∫T0(‖∇n‖2L2+‖Δc‖2L2+‖Δu‖2L2)dt≤C(‖n0‖2L2+‖∇c0‖2L2+‖∇u0‖2L2)exp(∫T01+‖∇u‖L∞+‖∇c‖2L∞). |
Observe that ‖n‖L∞(0,T;L2) and ‖∇n‖L2(0,T;L2) are uniformly bounded if ∫T0‖∇u‖L∞+‖∇c‖2L∞dt is bounded. Meanwhile, we notice that n∈LqxL∞t and ∇nq2∈L2xL2t for all 2<q<∞, and, owing to the following deduction
1qddt‖n‖qLq+Dn‖∇nq2‖2L2+μ∫R2nq+1dx≤C∫R2|nχ(c)∇c∇nq−1|dx+κ∫R2nqdx≤C‖∇c‖2L∞‖n‖qLq+12‖∇nq2‖2L2+κ‖n‖qLq. |
Collecting the above inequality, it ensures that ‖n(t)‖Lq≤C, with C is independent of q. Then, letting q→∞, n∈L∞xL∞t is obtained.
Additionally, we consider the estimate in the space (n,c,u)∈H1×H2×H2. It follows that
12ddt‖∇n‖2L2+Dn‖Δn‖2L2+2μ‖n12∇n‖2L2≤C‖∇(u⋅∇n)−(u⋅∇)∇n‖L2‖∇n‖L2+C‖χ(c)∇n∇c‖L2‖∇2n‖L2+C‖χ(c)nΔc‖L2‖∇2n‖L2+C‖χ′(c)n∇c∇c‖L2‖∇2n‖L2+κ‖∇n‖2L2≤C‖∇u‖L∞‖∇n‖2L2+C‖∇n‖L2‖∇c‖L∞‖∇2n‖L2+C‖n‖L∞‖Δc‖L2‖∇2n‖L2+C‖n‖L∞‖∇c‖L∞‖∇c‖L2‖∇2n‖L2+κ‖∇n‖2L2≤C‖∇u‖L∞‖∇n‖2L2+C‖∇n‖2L2‖∇c‖2L∞+18‖∇2n‖2L2+C‖n‖2L∞‖Δc‖2L2+18‖∇2n‖2L2+C‖n‖2L∞‖∇c‖2L∞‖∇c‖2L2+18‖∇2n‖2L2+κ‖∇n‖2L2. |
In conjunction with Young's inequality and Gronwall's inequality, we conclude
sup‖∇n‖2L2+∫T0‖∇2n‖2L2dt≤(‖∇n0‖2L2+C‖n‖2L∞(0,T;L∞)(∫T0‖Δc‖2L2dt+‖∇c‖2L∞(0,T;L2)∫T0‖∇c‖2L∞dt))×exp(∫T01+‖∇u‖L∞+‖∇c‖2L∞dt). |
Afterwards, n∈H1xL∞t∩H2xL2t. For the H2 estimate of c, we get
12ddt‖Δc‖2L2+Dc‖∇Δc‖2L2≤C‖∇u‖L∞‖Δc‖2L2+C‖Δu‖L2‖c‖L∞‖∇Δc‖L2+C‖g′(c)n∇c‖L2‖∇Δc‖L2+C‖g(c)n∇c‖L2‖∇Δc‖L2≤C‖∇u‖L∞‖Δc‖2L2+C‖Δu‖L2‖c‖L∞‖∇Δc‖L2+C‖∇c‖L2‖n‖L∞‖∇Δc‖L2+C‖∇n‖L2‖∇Δc‖L2≤C‖∇u‖L∞‖Δc‖2L2+C‖Δu‖2L2‖c‖2L∞+18‖∇Δc‖2L2+C‖∇c‖2L2‖n‖2L∞+18‖∇Δc‖2L2+C‖∇n‖2L2+18‖∇Δc‖2L2, |
and
12ddt‖Δu‖2L2+Du‖∇Δu‖2L2≤C‖∇u‖∞‖∇u‖2L2+C‖Δn‖L2‖Δu‖L2≤C‖∇u‖∞‖∇u‖2L2+C‖Δn‖2L2+12‖Δu‖2L2. |
Then we get c∈H2xL∞t∩H3xL2t and u∈H2xL∞t∩H3xL2t by Gronwall's inequality. In what follows, we show the estimate in the space (n,c,u)∈H2×H3×H3. In a similar way, it indicates
12ddt‖n‖2H2+Dn‖∇n‖2H2≤C‖u‖L∞‖n‖H2‖∇n‖H2+C‖∇u‖L∞‖n‖H1‖∇n‖H2C‖χ(c)n∇c‖H2‖∇n‖H2+κ‖n‖2H2+C‖n‖L∞‖n‖2H2≤C‖u‖2L∞‖n‖2H2+C‖∇u‖2L∞‖n‖2H1+C‖χ(c)n∇c‖2H2+κ‖n‖2H2+C‖n‖L∞‖n‖2H2+12‖∇n‖2H2. |
We control
‖χ(c)n∇c‖H2≤C‖n‖H2‖χ(c)∇c‖H2, |
and
‖∇2(χ(c)∇c)‖L2≤C‖∇3c‖L2+C‖∇2c‖L2‖∇c‖L∞+C‖∇c‖3L6. |
We also obtained c∈H2xL∞t∩H3xL2t. Consequently, in view of Young's inequality and Gronwall's inequality, it follows that
sup‖n‖2H2+∫T0‖∇n‖2H2dt≤‖n0‖2H2exp(C+∫T01+‖∇u‖L∞+‖∇c‖2L∞dt). |
In what follows, for the estimate of c, we further have
12ddt‖c‖2H3+Dc‖∇c‖2H3≤C‖∇u‖L∞‖c‖H3‖∇c‖H3+C‖u‖H3‖∇c‖L∞‖∇c‖H3+C‖g(c)n‖2H2+14‖∇c‖2H3≤C‖∇u‖2L∞‖c‖2H3+C‖u‖2H3‖∇c‖2L∞+C(‖c‖2H2‖n‖2H2+‖c‖2H1‖∇c‖2L∞‖n‖2H2)+12‖∇c‖2H3. |
For the estimate of u, we thereby obtain
12ddt‖u‖2H3+Du‖∇u‖2H3≤C‖∇u‖L∞‖u‖H3‖∇u‖H3+C‖n‖H2‖∇u‖H3≤C‖∇u‖2L∞‖u‖2H3+C‖n‖2H2+12‖∇u‖2H3. |
Applying Gronwall's inequality, it implies (c,u)∈(H3xL∞t∩H4xL2t)×(H3xL∞t∩H4xL2t). Let us demonstrate Hm−1×Hm×Hm estimates. We've proved that the case m=2,3 and 4, and therefore we consider the case m≥5. Operating ∂α(|α|≤m−1) and multiplying ∂αn on both sides of (1.1)1 and integrating in spaces, we thereby obtain
12ddt‖n‖2Hm−1+Dn‖∇n‖2Hm−1≤C‖∇u‖L∞‖n‖Hm−1‖∇n‖Hm−1+C‖u‖Hm−1‖∇n‖L∞‖∇n‖Hm−1+C‖χ(c)n∇c‖Hm−1‖∇n‖Hm−1+κ‖n‖2Hm−1+C‖n‖L∞‖n‖2Hm−1≤C‖∇u‖2L∞‖n‖2Hm−1+C‖u‖2Hm−1‖∇n‖2L∞+C‖χ(c)n∇c‖2Hm−1+κ‖n‖2Hm−1+C‖n‖L∞‖n‖2Hm−1+12‖∇n‖2Hm−1. |
The estimate for the case m=4 was already obtained, thus ‖∇c‖L∞(0,T;L∞) is bounded. Subsequently, we derive
‖∇χ(c)‖Hm−2≤C(1+‖∇c‖L∞)‖∇χ′(c)‖Hm−3. |
the classical product lemma on each step of iteration shows
‖∇χ(c)‖Hm−2≤C(1+‖∇c‖L∞)m−1. |
Therefore we also obtain
‖χ(c)n∇c‖Hm−1≤C(1+‖c‖Hm+‖∇c‖mL∞)‖n‖Hm−1 |
applying the product lemma. Add up the inequality above and we get
ddt‖n‖2Hm−1+‖∇n‖2Hm−1≤C‖∇u‖2L∞‖n‖2Hm−1+C‖u‖2Hm−1‖∇n‖2L∞+C(1+‖c‖Hm+‖∇c‖mL∞)2‖n‖2Hm−1+κ‖n‖2Hm−1+C‖n‖L∞‖n‖2Hm−1. |
Similarly, for the Hm estimate of c, we get
12ddt‖c‖2Hm+Dc‖∇c‖2Hm≤C‖∇u‖L∞‖c‖Hm‖∇c‖Hm+C‖u‖Hm‖∇c‖L∞‖∇c‖Hm+C(1+‖∇c‖2m−2L∞)‖n‖2Hm−1+14‖∇c‖2Hm≤C‖∇u‖2L∞‖c‖2Hm+C‖u‖2Hm‖∇c‖2L∞+C(1+‖∇c‖2m−2L∞)‖n‖2Hm−1+12‖∇c‖2Hm. |
For the estimate of u, we also deduce that
12ddt‖u‖2Hm+Du‖∇u‖2Hm≤C‖∇u‖L∞‖u‖Hm‖∇u‖Hm+C‖n‖Hm−1‖∇u‖Hm≤C‖∇u‖2L∞‖u‖2Hm+C‖n‖2Hm−1+12‖∇u‖2Hm. |
Finally, by summing up all the above estimates and utilizing Gronwall's inequality, the fact (n,c,u)∈(Hm−1xL∞t∩HmxL2t)×(HmxL∞t∩Hm+1xL2t)×(HmxL∞t∩Hm+1xL2t) is obtained.
As a matter of fact that ‖∇c‖L∞ is solely responsible for n∈L2xL∞t and ∇n∈L2xL2t by the above process. That is
ddt‖n‖2L2+‖∇n‖2L2+μ∫R2n3dx≤C∫R2|nχ(c)∇c∇n|dx+κ‖n‖2L2≤C‖∇c‖2L∞‖n‖2L2+12‖∇n‖2L2+κ‖n‖2L2. |
This yields u∈L2xL∞t and ∇u∈L2xL2t by
ddt‖u‖2L2+‖∇u‖2L2≤C‖n‖L2‖u‖L2≤C‖n‖2L2+C‖u‖2L2. |
Furthermore, we obtain n∈LqxL∞t and ∇nq2∈L2xL2t for all 2<q<∞
ddt‖n‖qLq+‖∇nq2‖2L2+μ∫R2nq+1dx≤Cq∫R2|nχ(c)∇c∇nq−1|dx+κ∫Rdnqdx≤Cq‖∇c‖2L∞‖n‖qLq+12‖∇nq2‖2L2+κ‖n‖qLq. |
Besides, we note that ∇c∈L2xL∞t and ∇2c∈L2xL2t. Actually,
ddt‖∇c‖2L2+‖∇2c‖2L2≤C‖∇c‖L∞‖u‖L2‖∇2c‖L2+C‖n‖L2‖∇2c‖L2≤C‖∇c‖2L∞‖u‖2L2+C‖n‖2L2+12‖∇2c‖2L2. |
Finally, we denote vorticity as ω:=∇×u; where ω=∂1u2−∂2u1 in two dimensions. Then, let us set ∇⊥n=(−∂2n,∂1n). We investigate the vorticity equation
ωt−Δω+u∇ω=∇⊥n∇Φ, |
We notice that ω∈L2xL∞t and ∇ω∈L2xL2t, on account of
ddt‖ω‖2L2+‖∇ω‖2L2≤C‖∇n‖L2‖ω‖L2≤C‖∇n‖2L2+C‖ω‖2L2. |
In addition, we note that ∇ω∈L2xL∞t and ∇2ω∈L2xL2t. In fact, testing −Δω, this implies that
ddt‖∇ω‖2L2+‖∇2ω‖2L2≤‖u‖L4‖∇ω‖L4‖Δω‖L2+‖∇n‖L2‖Δω‖L2≤C‖u‖12L2‖∇u‖12L2‖∇ω‖12L2‖∇2ω‖32L2+‖∇n‖L2‖Δω‖L2≤C‖u‖2L2‖∇u‖2L2‖∇ω‖2L2+C‖∇n‖2L2+12‖∇2ω‖2L2. |
Hence, via embedding, we get
∫T0‖∇u‖L∞dt≤∫T0‖∇u‖H2dt≤C∫T0‖ω‖H2dt<∞. |
This implies the desired result.
In this section, we will show that the local classical solutions can be extended at any time T>0.
Proof of Theorem 1.2. The process is similar to the proof of Theorem 1.2 in [7]. Using contradictory methods, supposing that the maximal time T∗ is finite, we will prove
∫T∗0‖∇c(τ)‖2L∞(R2)dτ<∞, |
which contradicts the extensibility criterion (1.10). Actually, according to the fact
‖∇c(τ)‖2L∞(R2)≤˜CGN‖∇c‖L2(R2)‖∇3c‖L2(R2)≤˜CGN2(‖∇c‖2L2(R2)+‖∇3c‖2L2(R2)), |
where ˜CGN is a positive constant resulted from the Gagliardo-Nirenberg inequality.
Now, we just need to verify that
∫T∗0∫R2|∇c(x,τ)|2dxdτ<∞, | (3.1) |
and
∫T∗0∫R2|∇3c(x,τ)|2dxdτ<∞. | (3.2) |
According to Proposition 1, we obtain (3.1).
As for (3.2), applying Δ to the (1.1)2, multiplying Δc with the resulted equation, and integrating over R2, we have
12ddt∫R2|Δc|2dx+Dc∫R2|∇Δc|2dx=∫R2(∇(u⋅∇c)+g′(c)n∇c+g(c)∇n)⋅∇Δcdx:=I1+I2+I3. |
For I1, we get
I1≤1Dc∫R2|∇u|2|∇c|2dx+1Dc∫R2|u|2|D2c|2dx+Dc4∫R2|∇Δc|2dx≤1Dc‖∇u‖2L3(R2)‖∇c‖2L6(R2)+1Dc‖u‖2L∞(R2)‖D2c‖2L2(R2)+Dc4∫R2|∇Δc|2dx. |
Similarly, let M=‖c0‖L∞, we conclude
I2≤1Dcsup0≤s≤Mg′2(s)∫R2|n|2|∇c|2dx+Dc8∫R2|∇Δc|2dx≤1Dcsup0≤s≤Mg′2(s)‖n‖2L3(R2)‖∇c‖2L6(R2)+Dc8∫R2|∇Δc|2dx, |
and
I3≤1Dcsup0≤s≤Mg2(s)∫R2|∇n|2dx+Dc8∫R2|∇Δc|2dx≤1Dcsup0≤s≤Mg2(s)∫R2|∇n|2dx+Dc8∫R2|∇Δc|2dx. |
Collecting I1−I3, we obtain
ddt∫R2|Δc|2dx+Dc∫R2|∇Δc|2dx≤C2(‖u‖2H2(R2)+‖n‖2H1(R2))‖∇c‖2H1(R2)+C2∫R2|∇n|2dx, |
with C2 is a positive constant depending on the Sobolev's embedding and the initial data. The Gronwall's inequality implies that
‖∇c(t)‖2H1(R2)+Dc∫T∗0‖∇2c(τ)‖2H1(R2)dτ≤‖c0‖2H2(R2)eC2∫T∗0(‖n(τ)‖2H1(R2)+‖u(τ)‖2H2(R2))dτ+C2∫T∗0‖n(τ)‖2H1(R2)dτ, |
for all t∈(0,T∗). Therefore we have verified (3.2) provided that
∫T∗0(‖n‖2H1(R2)+‖u‖2H1(R2))(τ)dτ<∞. |
According to Proposition 1, we verify the above inequality. For convenience of readers, we give a sketch of the proof. To begin with, taking the L2-inner product with n for equation (1.1)1, we get
12ddt∫R2n2dx+Dn∫R2|∇n|2dx+μ∫R2n3dx=∫R2nχ(c)∇c⋅∇ndx+κ∫R2n2dx≤Dn4∫R2|∇n|2dx+sup0≤s≤Mχ2(s)Dn∫R2n2|∇c|2dx+κ∫R2n2dx. |
Next, we control ∫R2n2|∇c|2dx of the above inequality
∫R2n2|∇c|2dx≤‖n‖2L4(R2)‖∇c‖2L4(R2)≤CGN‖n‖L2(R2)‖∇n‖2L2(R2)‖c‖L∞(R2)‖Δc‖L2(R2)≤Dn4∫R2|∇n|2dx+C3∫R2|Δc|2dx∫R2n2dx, |
with C3=sup0≤s≤Mχ4(s)C2GNM2D3n, by using Sobolev's embedding, Young's inequality and the boundedness of c. Therefore, we have
ddt∫R2n2dx+Dn∫R2|∇n|2dx≤(2C3∫R2|Δc|2dx+κ)∫R2n2dx. |
In view of Gronwall's inequality, we deduce
∫R2n2(x,t)dx+Dn∫T∗0∫R2|∇n(x,τ)|2dx≤C4, |
for all t∈(0,T∗) and some positive constant C4 depending on the initial data and the maximal time T∗, where we also used the inequality
∫T∗0∫R2|Δc(x,τ)|2dxdτ<∞, |
by recalling Proposition 1.1. A direct integration on [0,T∗] implies that
∫T∗0‖n(τ)‖2H1(R2)dτ<∞. | (3.3) |
Similarly, we also have
∫T∗0‖u(τ)‖2H1(R2)dτ<∞. | (3.4) |
Let's first investigate the integrability of the second derivative of u. For convenience, let ω:=∇⊥u be the vorticity of u and then the vorticity equation as follow:
ωt+(u⋅∇)ω=DuΔω+∇⊥(n∇Φ). |
Next, a direct energy method follows that
12ddt∫R2|ω|2dx+Du∫R2|∇ω|2dx=∫R2∇⊥ω⋅(n∇Φ)dx≤Du2∫R2|∇ω|2dx+12Du∫R2n2|∇Φ|2dx. |
This leads to the vorticity estimate
ddt∫R2|ω|2dx+Du∫R2|∇ω|2dx≤C5∫R2n2dx, |
where C5 is a positive constant depending on ∇Φ. In conjunction with Gronwall's inequality and (3.3), we infer
∫R2|ω|2(x,t)dx+Du∫T∗0∫R2|∇ω(x,τ)|2dxdτ≤C6, |
for all t∈(0,T∗) with C6:=∫R2|ω|2(x,0)dx+C4C5T∗. Application of the above inequality and the Biot-Savart law, we derive
∫T∗0‖∇2u(τ)‖2L2(R2)dτ<∞. | (3.5) |
Sum up (3.3)-(3.5), it implies the desired result. This completes the Proof of Theorem 1.2.
Q. Zhang was partially supported by the Natural Science Foundation of Hebei Province [grant number A2020201014 and A2019201106]; the Second Batch of Young Talents of Hebei Province; Nonlinear Analysis Innovation Team of Hebei University.
All authors declare that there is no interests in this paper.
In this appendix, we will give a sketch proof of local existence.
We first introduce the dynamic partition of the unity to define Besov spaces. One may check [15] for more details. Let φ∈C∞0(Rd) be supported in C={ξ∈Rd,34≤|ξ|≤83} such that
∑q∈Zφ(2−qξ)=1,forξ≠0. |
Defining χ(ξ)=1−∑q∈Nφ(2−qξ). For f∈S′, we set Littlewood-Paley operators as follows
Δ−1f=χ(D)f;∀q∈N,Δqf=φ(2−qD)fand∀q∈Z,˙Δqf=φ(2−qD)f. |
The following low-frequency cut-off will be also used:
Sqf=∑−1≤q′≤q−1Δq′fand˙Sqf=∑q′≤q−1˙Δq′f. |
Now, let us recall the definition of the Besov space. For s∈R, 1≤p,r≤∞, we define the homogenous Besov space ˙Bsp,r as the set of tempered distributions of f∈S′/P such that
‖f‖˙Bsp,r:=(∑q∈Z2qsr‖˙Δqf‖rp)1r<∞, |
where P is the polynomial space. The inhomogeneous space Bsp,r is the set of tempered distribution f such that
‖f‖Bsp,r:=(∑q≥−12qsr‖Δqf‖rp)1r<∞. |
It is worthwhile to remark that Bs2,2 and Bs∞,∞ coincide with the usual Sobolev spaces Hs and the usual Hölder space Cs for s∈R∖Z, respectively.
In our study, we require the space-time Besov spaces as following manner: for T>0 and n≥1, we denote by LρTBsp,r the set of all tempered distribution f satisfying
‖f‖LρTBsp,r≜‖(∑q∈Z2qsr‖˙Δqf‖rLp(Rd))1r‖LρT<∞. |
Lemma A.1. [15] Let 1≤p≤q≤∞. Assume that f∈Lp, then there exists a constant C independent of f, j such that
suppˆf⊂{|ξ|≤C2j}⟹‖∂αf‖q≤C2j|α|+dj(1p−1q)‖f‖p, |
suppˆf⊂{1C2j≤|ξ|≤C2j}⟹‖f‖p≤C2−j|α|sup|β|=|α|‖∂βf‖p. |
Lemma A.2. [15] There exists a constant C>0 such that for S>0, we have
‖uv‖Hs≤C‖u‖∞‖v‖Hs+C‖u‖Hs‖v‖∞. |
Lemma A.3. [5] Let u be a solution of the transport equation
{ut+v⋅∇u=0u(x,0)=u0 |
and define Rq:=v⋅∇Δqu−Δq(v⋅∇u), 1≤p≤p1≤∞, 1≤r≤∞ and s∈R such that
s>−dmin(1p1,1p′)(ors>−1−dmin(1p1,1p′)ifdivv=0). |
There exists a sequence cq∈ℓr(Z) such that ‖cq‖ℓr=1 and a constant C depending only on d,r,s,p, and p1, which satisfy
∀q∈Z,2qs‖Rq‖p≤CcqZ′(t)‖u‖Bsp,r, |
with
Z′(t):={‖∇v‖Bdp1p1,∞∩L∞,ifs<1+dp1,‖∇v‖Bs−1p1,r,ifeithers>1+dp1ors=1+dp1forr=1. |
Theorem A.4. Let s≥2. Assume that χ(⋅), g(⋅)∈Cs(R) with g(0)=0, and that ‖∇lΦ‖L∞(R2)<∞ for 1≤|l|≤s.
(A1) Then there exists T∗>0, the maximal time of existence, such that, if the initial data (n0,c0,u0)∈Hs−1(R2)×Hs(R2)×Hs(R2;R2), then there exists a unique classical solution (n,c,u) to system (1.1)-(1.2) satisfying for any T<T∗
(n,c,u)∈L∞(0,T;Hs(R2)×Hs+1(R2)×Hs+1(R2;R2)) |
and
(∇n,∇c,∇u)∈L2(0,T;Hs(R2)×Hs+1(R2)×Hs+1(R2;R2)). |
Proof. We construct the following regularized system:
{nkt+uk⋅∇nk=Δnk−∇⋅(nkχ(ck)∇ck)+nk−(nk)2,k∈N,ckt+uk⋅∇ck=Δck−g(ck)nk,ukt+(uk⋅∇)uk−∇Pk=Δuk+nk∇Φ,∇⋅uk=0,(nk,ck,uk)|t=0=(Skn0,Skc0,Sku0). | (A.1) |
Step 1. Uniform estimates.
Taking the operation Δq with q≥−1 on the first equation of (A.1), we have
Δqnkt+Δq(uk⋅∇nk)=ΔΔqnk−∇⋅Δq(nkχ(ck)∇ck)+Δqnk−Δq(nk)2. | (A.2) |
Multiplying (A.2) by Δqnk and integrating by parts gives
12ddt‖Δqnk‖22+‖∇Δqnk‖22=−∫R2Δq(uk⋅∇nk)Δqnkdx−∫R2∇⋅Δq(nkχ(ck)∇ck)Δqnkdx+∫R2ΔqnkΔqnkdx−∫R2Δq(nk)2Δqnkdx≤‖Δq(uk⋅∇nk)‖2‖Δqnk‖2+‖Δq(nk∇ck)‖2‖∇Δqnk‖2+‖Δqnk‖22+‖Δq(nk)2‖2‖Δqnk‖2. |
Multiplying 22qs on both sides of the above inequality, then taking the ℓ1 norm, using Hölder's inequality and Young's inequality together with Lemma 4.2, we get
12ddt‖nk‖2Hs+‖nk‖2Hs+1≤‖uk⋅∇nk‖Hs‖nk‖Hs+‖nk∇ck‖Hs‖nk‖Hs+1+‖nk‖2Hs+‖(nk)2‖Hs‖nk‖Hs≤C‖uk‖Hs‖nk‖Hs+1‖nk‖Hs+C‖nk‖Hs‖ck‖Hs+1‖nk‖Hs+1+‖nk‖2Hs+C‖nk‖2Hs‖nk‖Hs≤C‖uk‖2Hs‖nk‖2Hs+18‖nk‖2Hs+1+C‖nk‖2Hs‖ck‖2Hs+1+18‖nk‖2Hs+1+‖nk‖2Hs+C(‖nk‖4Hs+‖nk‖2Hs). |
from which we have
ddt‖nk‖2Hs+‖nk‖2Hs+1≤C(‖uk‖2Hs‖nk‖2Hs+‖nk‖2Hs‖ck‖2Hs+1+‖nk‖2Hs+‖nk‖4Hs). | (A.3) |
In a similarly way to (A.3), we obtain
12ddt‖ck‖2Hs+1+‖ck‖2Hs+2≤C‖uk‖2Hs+1‖ck‖2Hs+1+18‖ck‖2Hs+2+C‖ck‖2Hs+1+18‖nk‖2Hs+1. |
it follows that
ddt‖ck‖2Hs+1+‖ck‖2Hs+2≤C(‖uk‖2Hs+1‖ck‖2Hs+1+‖ck‖2Hs+1)+18‖nk‖2Hs+1. | (A.4) |
Applying Δq with q≥−1 to the third equation of (A.1) yields
Δqukt+(uk⋅∇)Δquk−∇ΔqPk=(uk⋅∇)Δquk−Δq((uk⋅∇)uk)+Δq(nk∇Φ). |
Multiplying the above equality with Δquk yields
12ddt‖Δquk‖22+‖∇Δquk‖22=∫Rd((uk⋅∇)Δquk−Δq((uk⋅∇)uk))Δqukdx+∫RdΔq(nk∇Φ)Δqukdx≤‖ (uk⋅∇)Δquk−Δq((uk⋅∇)uk)‖2‖Δquk‖2+‖Δq(nk∇Φ)‖2‖Δquk‖2. |
Multiplying 22q(s+1) on both side of the above inequality and taking the l1 norm, we obtain
ddt‖uk‖2Hs+1+‖uk‖2Hs+2≤‖∇uk‖∞‖uk‖2Hs+1+18‖nk‖2Hs+1+‖uk‖2Hs+1≤C(‖uk‖4Hs+1+‖uk‖2Hs+1)+18‖nk‖2Hs+1. | (A.5) |
Summing up (A.3)-(A.5), we obtain
ddt(‖nk‖2Hs+‖ck‖2Hs+1+‖uk‖2Hs+1)+‖nk‖2Hs+1+‖ck‖2Hs+2+‖uk‖2Hs+2≤C(‖nk‖2Hs+‖ck‖2Hs+1+‖uk‖2Hs+1)(1+‖nk‖2Hs+‖ck‖2Hs+1+‖uk‖2Hs+1)≤(1+‖nk‖2Hs+‖ck‖2Hs+1+‖uk‖2Hs+1)2. |
We conclude from the Gronwall inequality that
1+‖nk‖2Hs+‖ck‖2Hs+1+‖uk‖2Hs+1≤1+‖nk0‖2Hs+‖ck0‖2Hs+1+‖uk0‖2Hs+11−C(1+‖nk0‖2Hs+‖ck0‖2Hs+1+‖uk0‖2Hs+1)t. |
We can choose
T=12C(1+‖nk0‖2Hs+‖ck0‖2Hs+1+‖uk0‖2Hs+1)>0. |
such that
supt∈[0,T](‖nk(t)‖2Hs+‖ck(t)‖2Hs+1+‖uk(t)‖2Hs+1)+∫t0(‖nk‖2Hs+1+‖ck‖2Hs+2+‖uk‖2Hs+2)(τ)dτ≤2(1+‖nk0‖2Hs+‖ck0‖2Hs+1+‖uk0‖2Hs+1). | (A.6) |
Step 2. Compactness.
From (A.6), we obtain
nk∈L∞([0,T],Hs)∩L2([0,T],Hs+1),ck∈L∞([0,T],Hs+1)∩L2([0,T],Hs+2),uk∈L∞([0,T],Hs+1)∩L2([0,T],Hs+2). |
In order to show the convergence, we also need uniform boundedness for ∂tnk, ∂tck and ∂tuk. From the first equation of (A.1), we know
‖∂tnk‖L∞tH−1≤‖Δnk‖L∞tH−1+‖uk⋅∇nk‖L∞tH−1+‖∇⋅(nk∇ck)‖L∞tH−1+‖nk‖L∞tH−1+‖(nk)2‖L∞tH−1≤‖nk‖L∞tHs+‖uk‖L∞tHs+1‖nk‖L∞tHs+‖nk‖L∞tHs‖ck‖L∞tHs+1+‖nk‖L∞tHs+‖nk‖2L∞tHs≤C. |
Similarly, we have
‖∂tck‖L∞tH−1≤‖Δck‖L∞tH−1+‖uk⋅∇ck‖L∞tH−1+‖cknk‖L∞tH−1≤‖ck‖L∞tHs+1+‖uk‖L∞tHs+1‖ck‖L∞tHs+1+‖ck‖L∞tHs+1‖nk‖L∞tHs≤C. |
and
‖∂tuk‖L∞tH−1≤‖Δuk‖L∞tH−1+‖(uk⋅∇)uk‖L∞tH−1+‖nk∇Φ‖L∞tH−1≤‖uk‖L∞tHs+1+‖uk‖2L∞tHs+1+‖nk‖L∞tHs≤C. |
Since L2 is locally compactly embedded in H−1, we apply the Aubin-Lions Lemma to conclude that, up to an extraction of subsequence, the approximate solution sequence (nk,ck,uk) strongly converges in L∞([0,T];H−1) to some function (n,c,u) such that
nk∈L∞([0,T];Hs)∩L2([0,T],Hs+1),ck∈L∞([0,T];Hs+1)∩L2([0,T],Hs+2),uk∈L∞([0,T];Hs+1)∩L2([0,T],Hs+2). |
Using the above estimates, it is easy to pass the limit in the approximate system (A.1) and (n,c,u) solve (1.1) in the sense of distribution. By a classical deduction [34], we get n∈C([0,T];Hs), c∈C([0,T];Hs+1) and u∈C([0,T];Hs+1).
By virtue of Heat equation theory, we can prove the time differentiation. For example, we consider the following equations:
ut−Δu=f(x,t),thenut=Δu−f(x,t). |
Suppose u∈L∞tHs and f∈L∞tHs−2, we obtain ut∈L∞tHs−2. For the arbitrariness of s, we have ut∈L∞tHs,∀ s>0. In a similar way, we get
∂t(ut)−Δut=ft |
and utt∈L∞tHs−2. Thus we show the time differentiation.
Step 3. Uniqueness.
Let us consider the two solutions (n1,c1,u1) and (n2,c2,u2) associated with the same initial data and satisfy (1.1). We use the notation δn=n1−n2, δc=c1−c2 and δu=u1−u2. Then we have
{∂tδn+δu⋅∇n1+u2⋅∇δn=Δδn−∇⋅(δnχ(c1)∇c1)−∇⋅(n2(χ(c1)−χ(c2))∇c1)−n2χ(c2)∇(δc)+δn−n1δn−n2δn,∂tδc+δu⋅∇c1+u2⋅∇δc=Δδc−n1(g(c1)−g(c2))−g(c2)δn,∂tδu+(δu⋅∇)u1+(u2⋅∇)δu−∇(P1−P2)=Δδu+δn∇Φ. | (A.7) |
Taking the L2-inner product of the first equation with δn, we have
12ddt‖δn(t)‖22+‖∇δn(t)‖22=−∫R2(δu⋅∇n1)δndx−∫R2∇⋅(δnχ(c1)∇c1)δndx−∫R2∇⋅(n2(χ(c1)−χ(c2))∇c1)δndx−∫R2n2χ(c2)∇δcδndx+∫R2δnδndx−∫R2n1δnδndx−∫R2n2δnδndx≤C(‖δu‖22+‖δn‖22‖n1‖2Hs)+C‖δn‖22‖c1‖2Hs+1+C‖δc‖22‖n2‖2Hs‖c1‖2Hs+1+18‖∇δn‖22+C‖n2‖2Hs‖∇δc‖22+C‖δn‖22+‖n1‖Hs‖δn‖22+‖n2‖Hs‖δn‖22. |
from which we get
ddt‖δn(t)‖22+‖∇δn(t)‖22≤C(‖δu‖22+‖δn‖22‖n1‖2Hs+‖δn‖22‖c1‖2Hs+1+‖n2‖2Hs‖∇δc‖22+‖δn‖22+‖n1‖Hs‖δn‖22+‖n2‖Hs‖δn‖22+‖δc‖22‖n2‖2Hs‖c1‖2Hs+1). | (A.8) |
Next, taking the L2-inner product of the second equation of Eqs.(A.7) with δc, we know
12ddt‖δc(t)‖22+‖∇δc(t)‖22=−∫R2(δu⋅∇c1)δcdx−∫R2(g(c1)−g(c2))n1δcdx−∫R2δng(c2)δcdx≤C(‖δu‖22+‖δc‖22‖c1‖2Hs+1+‖δc‖22‖n1‖Hs+‖δn‖22+‖δc‖22). |
Hence we get
ddt‖δc(t)‖22+‖∇δc(t)‖22≤C(‖δu‖22+‖δc‖22‖c1‖2Hs+1+‖δc‖22‖n1‖Hs+‖δn‖22+‖δc‖22). | (A.9) |
Taking ∂i on both sides of the second equation of Eqs (A.7) yields
∂t∂iδc+u2⋅∇∂i∇c−Δ∂iδc=−∂i(δu⋅∇c1)−∂iu2⋅∇δc−∂i(n1(g(c1)−g(c2)))−∂i(g(c2)δn). |
Multiplying the above equation with ∂iδc and integrating with respect to space variable, we obtain
12ddt‖∇δc(t)‖22+‖Δδc(t)‖22=−∑i∫Rd∂i(δu⋅∇c1)∂iδcdx−∑i∫Rd∂iu2⋅∇δc∂iδcdx−∑i∫Rd∂i(n1(g(c1)−g(c2)))∂iδcdx−∑i∫Rd∂i(g(c2)δn)∂iδcdx≤∫Rd(δu⋅∇c1)Δδcdx−∫Rd(∇δc⋅∇)u2⋅∇δcdx+C∫Rdn1δcΔδcdx+∫Rdg(c2)δnΔδcdx≤C‖δu‖22‖c1‖2Hs+1+18‖Δδc‖22+C‖∇δc‖22‖u2‖Hs+1+C‖δc‖22‖n1‖2Hs+18‖Δδc‖22+C‖δn‖22+18‖Δδc‖22. |
Then we get
ddt‖∇δc(t)‖22+‖Δδc(t)‖22≤C(‖δu‖22‖c1‖2Hs+1+‖∇δc‖22‖u2‖Hs+1+‖δc‖22‖n1‖2Hs+‖δn‖22). | (A.10) |
Performing the L2-inner product of the third equation of system (A.7) with δu, we get
12ddt‖δu(t)‖22+‖∇δu(t)‖22=−∫Rd((δu⋅∇)u1)⋅δudx+∫Rdδn∇Φ⋅δudx≤C‖δu‖22‖u1‖Hs+1+C(‖δn‖22+‖δu‖22). |
Such that
ddt‖δu(t)‖22≤C(‖δu‖22‖u1‖Hs+1+‖δn‖22+‖δu‖22). | (A.11) |
From ((A.8)-(A.11), we have
ddt(‖δn(t)‖22+‖δc(t)‖22+‖∇δc(t)‖22+‖δu(t)‖22)+‖∇δn‖22+‖∇δc‖22+‖Δδc‖22+‖∇δu(t)‖22≤C(‖δu‖22+‖δn‖22‖n1‖2Hs+‖δn‖22‖c1‖2Hs+1+‖n2‖2Hs‖∇δc‖22+‖δn‖22+‖n1‖Hs‖δn‖22+‖n2‖Hs‖δn‖22+‖δc‖22‖n2‖2Hs‖c1‖2Hs+1+‖δu‖22+‖δc‖22‖c1‖2Hs+1+‖δc‖22‖n1‖Hs+‖δn‖22+‖δc‖22+‖δu‖22‖c1‖2Hs+1+‖∇δc‖22‖u2‖Hs+1+‖δc‖22‖n1‖2Hs+‖δn‖22+‖δu‖22‖u1‖Hs+1). |
Consequently, we have
ddt(‖δn(t)‖22+‖δc(t)‖22+‖∇δc(t)‖22+‖δu(t)‖22)≤CF(t)(‖δn‖22+‖δc‖22+‖∇δc‖22+‖δu‖22) |
where
F(t)=1+‖n1‖Hs+‖n2‖Hs+‖u1‖Hs+1+‖u2‖Hs+1+‖n1‖2Hs+‖n2‖2Hs+‖c1‖2Hs+1+‖c2‖2Hs+1+‖n2‖2Hs‖c1‖2Hs+1. |
By the above estimates, we know that F(t) is integrable. Therefore, we finally obtain the uniqueness by using the Gronwall inequality.
[1] |
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
![]() |
[2] |
X. Cao, J. Lankeit, Global classical small-data solutions for a three-dimensional chemotaxis Navier-Stokes system involving matrix-valued sensitivities, Calc. Var. Partial Dif., 55 (2016), 107. https://doi.org/10.1007/s00526-016-1027-2 doi: 10.1007/s00526-016-1027-2
![]() |
[3] |
M. Chae, K. Kang, J. Lee, On existence of the smooth solutions to the coupled chemotaxis-fluid equations, Discrete Contin. Dyn. Syst., 33 (2013), 2271–2297. https://doi.org/10.3934/dcds.2013.33.2271 doi: 10.3934/dcds.2013.33.2271
![]() |
[4] |
M. Chae, K. Kang, J. Lee, Global existence and temporal decay in Keller-Segel models coupled to fluid equations, Commun. Part. Diff. Eq., 39 (2014), 1205–1235. https://doi.org/10.1080/03605302.2013.852224 doi: 10.1080/03605302.2013.852224
![]() |
[5] | R. Danchin, A few remarks on the Camassa-Holm equation, Differ. Integral Equ., 14 (2001), 953–988 |
[6] |
R. Duan, A. Lorz, P. Markowich, Global solutions to the coupled chemotaxis-fluid equations, Commun. Part. Diff. Eq., 35 (2010), 1635–1673. https://doi.org/10.1080/03605302.2010.497199 doi: 10.1080/03605302.2010.497199
![]() |
[7] |
R. Duan, X. Li, Z. Xiang, Global existence and large time behavior for a two dimensional chemotaxis-Navier-Stokes system, J. Differential Equations 263 (2017), 6284–6316. https://doi.org/10.1016/j.jde.2017.07.015 doi: 10.1016/j.jde.2017.07.015
![]() |
[8] |
R. Duan, Z. Xiang, A note on global existence for the chemotaxis Stokes model with nonlinear diffusion, Int. Math. Res. Not. IMRN, 2014 (2014), 1833–1852. https://doi.org/10.1093/imrn/rns270 doi: 10.1093/imrn/rns270
![]() |
[9] |
K. Fujie, M. Winkler, T. Yokota, Blow-up prevention by logistic sources in a parabolic-elliptic Keller-Segel system with singular sensitivity, Nonlinear Anal., 109 (2014), 56–71. https://doi.org/10.1016/j.na.2014.06.017 doi: 10.1016/j.na.2014.06.017
![]() |
[10] |
X. He, S. Zheng, Convergence rate estimates of solutions in a higher dimensional chemotaxis system with logistic source, J. Math. Anal. Appl., 436 (2) (2016), 970–982. https://doi.org/10.1016/j.jmaa.2015.12.058 doi: 10.1016/j.jmaa.2015.12.058
![]() |
[11] |
J. Lankeit, Chemotaxis can prevent thresholds on population density, Discrete Contin. Dyn. Syst. Ser. B, 20 (2015), 1499–1527. https://doi.org/10.3934/dcdsb.2015.20.1499 doi: 10.3934/dcdsb.2015.20.1499
![]() |
[12] |
J. Lankeit, Long-term behaviour in a chemotaxis-fluid system with logistic source, Math. Models Methods Appl. Sci., 26 (2016), 2071–2109. https://doi.org/10.1142/S021820251640008X doi: 10.1142/S021820251640008X
![]() |
[13] |
J. G. Liu, A. Lorz, A coupled chemotaxis-fluid model: global existence, Ann. Inst. H. Poincaré. Anal. Non linéare, 28 (2011), 643–652. https://doi.org/10.1016/j.anihpc.2011.04.005 doi: 10.1016/j.anihpc.2011.04.005
![]() |
[14] | Y. Lin, Q. Zhang, Global well-posedness for the 2D chemotaxisfluid system with logistic source, Appl. Anal., (2020). https://doi.org/10.1080/00036811.2020.1767287 |
[15] | C. Miao, J. Wu, Z. Zhang, Littlewood-Paley Theory and Applications to Fluid Dynamics Equations, Monographs on Modern Pure Mathematics, No. 142, Science Press, Beijing, 2012. |
[16] |
Y. Tao, M. Winkler, Global existence and boundedness in a Keller-Segel-Stokes model with arbitrary porous medium diffusion, Discrete Contin. Dyn. Syst., 32 (2012), 1901–1914. https://doi.org/10.3934/dcds.2012.32.1901 doi: 10.3934/dcds.2012.32.1901
![]() |
[17] |
Y. Tao, M. Winkler, Boundedness and decay enforced by quadratic degradation in a three-dimensional chemotaxis-fluid system, Z. Angew. Math. Phys., 66 (2015), 2555–2573. https://doi.org/10.1007/s00033-015-0541-y doi: 10.1007/s00033-015-0541-y
![]() |
[18] | Y. Tao, M. Winkler, Blow-up prevention by quadratic degradation in a two-dimensional Keller-Segel- Navier-Stokes system, Z. Angew. Math. Phys., 67 (2016), 138. |
[19] |
I. Tuval, L. Cisneros, C. Dombrowski, C. W. Wolgemuth, J. O. Kessler, R. E. Goldstein, Bacterial swimming and oxygen transport near contact lines, Proc. Natl. Acad. Sci. USA, 102 (2005), 2277–2282. https://doi.org/10.1073/pnas.0406724102 doi: 10.1073/pnas.0406724102
![]() |
[20] | Y. Wang, M. Winkler, Z. Xiang, Global classical solutions in a two-dimensional chemotaxis-Navier-Stokes system with subcritical sensitivity, Ann. Sc. Norm. Super. Pisa Cl. Sci., 5 (2018). https://doi.org/10.2422/2036-2145.201603_004 |
[21] |
Y. Wang, Z. Xiang, Global existence and boundedness in a Keller-Segel-Stokes system involving a tensor-valued sensitivity with saturation, J. Differential Equations, 259 (2015), 7578–7609. https://doi.org/10.1016/j.jde.2015.08.027 doi: 10.1016/j.jde.2015.08.027
![]() |
[22] |
M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel, model, J. Differential Equations, 248 (2010), 2889–2905. https://doi.org/10.1016/j.jde.2010.02.008 doi: 10.1016/j.jde.2010.02.008
![]() |
[23] |
M. Winkler, Finite-time blow-up in the higher-dimensional parabolic-parabolic Keller-Segel, system, J. Math. Pures Appl., 100 (2013), 748–767. https://doi.org/10.1016/j.matpur.2013.01.020 doi: 10.1016/j.matpur.2013.01.020
![]() |
[24] |
M. Winkler, Emergence of large population densities despite logistic growth restrictions in fully parabolic chemotaxis systems, Discrete Contin. Dyn. Syst. Ser. B, 22 (2017), 2777–2793. https://doi.org/10.3934/dcdsb.2017135 doi: 10.3934/dcdsb.2017135
![]() |
[25] |
M. Winkler, Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source, Commun. Part. Diff. Eq., 35 (2010), 1516–1537. https://doi.org/10.1080/03605300903473426 doi: 10.1080/03605300903473426
![]() |
[26] |
M. Winkler, Global large-data solutions in a chemotaxis-(Navier-)Stokes system modeling cellular swimming in fluid drops, Commun. Part. Diff. Eq., 37 (2012), 319–351. https://doi.org/10.1080/03605302.2011.591865 doi: 10.1080/03605302.2011.591865
![]() |
[27] |
M. Winkler, Stabilization in a two-dimensional chemotaxis-Navier–Stokes system, Arch. Ration. Mech. Anal., 211 (2014), 455–487. https://doi.org/10.1007/s00205-013-0678-9 doi: 10.1007/s00205-013-0678-9
![]() |
[28] |
M. Winkler, Global weak solutions in a three-dimensional chemotaxis-Navier-Stokes system, Ann. Inst. H. Poincaré Anal. Non Linéaire, 33 (2016), 1329–1352. https://doi.org/10.1016/j.anihpc.2015.05.002 doi: 10.1016/j.anihpc.2015.05.002
![]() |
[29] |
M. Winkler, How far do chemotaxis-driven forces influence regularity in the Navier-Stokes system?, Trans. Amer. Math. Soc., 369 (2017), 3067–3125. https://doi.org/10.1090/tran/6733 doi: 10.1090/tran/6733
![]() |
[30] |
M. Winkler, A three-dimensional Keller-Segel-Navier-Stokes system with logistic source: Global weak solutions and asymptotic stabilization, J. Funct. Anal., 276 (2019), 1339–1401. https://doi.org/10.1016/j.jfa.2018.12.009 doi: 10.1016/j.jfa.2018.12.009
![]() |
[31] |
M. Winkler, Reaction-driven relaxation in three-dimensional Keller-Segel-Navier-Stokes interaction, Commun. Math. Phys., 389 (2022), 439–489. https://doi.org/10.1007/s00220-021-04272-y doi: 10.1007/s00220-021-04272-y
![]() |
[32] |
J. Wu and C. Wu, A note on the global existence of a two-dimensional chemotaxis-Navier-Stokes system, Appl. Anal., 98 (2019), 1224–1235. https://doi.org/10.1080/00036811.2017.1419199 doi: 10.1080/00036811.2017.1419199
![]() |
[33] |
Q. Zhang, Blowup criterion of smooth solutions for the incompressible chemotaxis-Euler equations, Z. Angew. Math. Mech., 96 (2016), 466–476. https://doi.org/10.1002/zamm.201500040 doi: 10.1002/zamm.201500040
![]() |
[34] |
Q. Zhang, X. Zheng, Global well-posedness for the two-dimensional incompressible chemptaxis-Navier-Stokes equations, SIAM J. Math. Anal., 46 (2014), 3078–3105. https://doi.org/10.1137/130936920 doi: 10.1137/130936920
![]() |
[35] |
X. Zhao, S. Zheng, Global boundedness to a chemotaxis system with singular sensitivity and logistic source, Z. Angew. Math. Phys., 68 (2017), 2. https://doi.org/10.1007/s00033-016-0749-5 doi: 10.1007/s00033-016-0749-5
![]() |
1. | Jijie Zhao, Xiaoyu Chen, Qian Zhang, Global Existence of Weak Solutions for the 2D Incompressible Keller-Segel-Navier-Stokes Equations with Partial Diffusion, 2022, 181, 0167-8019, 10.1007/s10440-022-00529-3 |