Next Article in Journal
Hölder-Type Inequalities for Power Series of Operators in Hilbert Spaces
Previous Article in Journal
An Efficient Subspace Minimization Conjugate Gradient Method for Solving Nonlinear Monotone Equations with Convex Constraints
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Global Existence of Chemotaxis-Navier–Stokes System with Logistic Source on the Whole Space R2

1
Department of Mathematics, Nanchang Normal University, Nanchang 330029, China
2
School of Statistics, Jiangxi University of Finance and Economics, Nanchang 330032, China
3
Department of Mathematics, Jiangxi University of Finance and Economics, Nanchang 330032, China
*
Author to whom correspondence should be addressed.
Axioms 2024, 13(3), 171; https://doi.org/10.3390/axioms13030171
Submission received: 13 January 2024 / Revised: 25 February 2024 / Accepted: 27 February 2024 / Published: 6 March 2024
(This article belongs to the Topic Advances in Nonlinear Dynamics: Methods and Applications)

Abstract

:
In this article, we study the Cauchy problem of the chemotaxis-Navier–Stokes system with the consumption and production of chemosignals with a logistic source. The parameters χ 0 ,   ξ 0 ,   λ > 0 and μ > 0 . The system is a model that involves double chemosignals; one is an attractant consumed by the cells themselves, and the other is an attractant or a repellent produced by the cells themselves. We prove the global-in-time existence and uniqueness of the weak solution to the system for a large class of initial data on the whole space R 2 .
MSC:
Primary: 42B30; Secondary: 35Q30; 42B35; 35B40

1. Introduction

The present paper is concerned with the following chemotaxis-Navier–Stokes system with the consumption and production of chemosignals with logistic source
u t + κ ( u · ) u = Δ u + P + n ϕ , · u = 0 , n t + ( u · ) n = Δ n χ · ( n c ) + ξ · ( n v ) + λ n μ n 2 , v t + ( u · ) v = Δ v v + n , c t + u · c = Δ c n c ,
where x R 2 , t > 0 . The terms n = n ( x , t ) ,   c = c ( x , t ) ,   v = v ( x , t ) ,   u = u ( x , t ) = ( u 1 ( x , t ) , u 2 ( x , t ) ) and P denote the unknown density of amoebae, the unknown oxygen concentration, the unknown concentration of the chemical attractant, and the unknown fluid velocity field and the unknown pressure, respectively. The parameters χ 0 ,   ξ 0 and κ 0 . The terms λ 0 and μ 0 reflect the rate of reproduction and death, respectively. We impose the following intial data
( u ( x , 0 ) , n ( x , 0 ) , v ( x , 0 ) , c ( x , 0 ) ) = ( u 0 , n 0 , v 0 , c 0 ) .
The time-independent function ϕ = ϕ ( x ) denotes the potential function produced by different physical mechanisms, e.g., the gravitational force or centrifugal force.
In some biological processes, chemotactic cells often interact with multiple chemotactic cues, either of which may be an attractant or a repellant, to produce variety of intricate patterns. It was pointed out in [1,2,3,4] that this phenomenon is widely present in many prototypical biological situations. As compared to the chemotaxis-fluid models involving just one chemical signal that is consumed or produced by the species themselves as mentioned above, chemotaxis-fluid models incorporating at least two different chemical signals seem much less understood. When χ > 0 ,   ξ < 0 ,   λ = μ = 0 and κ = 1 , the system (1) becomes an attraction–attraction Navier–Stokes system, and the corresponding Cauchy problem admits global mild solutions with small initial data in some scaling invariant space [5]. When χ > 0 ,   ξ > 0 and λ = κ = μ = 0 , the system (1) becomes an attraction-repulsion-Stokes system; the global bounded classical solution and the large time behavior of the solution have been established in a smoothly bounded planar domain [6,7]. When κ = 1 and χ > 0 ,   ξ > 0 ,   λ = μ = 0 , in [8] the corresponding attraction-repulsion Navier–Stokes system is proved to possess a unique global classical solution; however, the uniform boundedness and large time behavior of the solutions to this attraction-repulsion Navier–Stokes system can be achieved simultaneously in [9]. We refer to [10,11,12,13,14,15,16,17,18,19,20,21,22] for more details concerning some properties of chemotaxis-fluid models.
In this work, we shall focus on the Cauchy problem (1) with the logistic term in two dimensional. Here, the parameters χ 0 ,   ξ 0 ,   λ > 0 ,   μ > 0 and κ = 1 . Precisely, we shall consider the global-in-time existence and uniqueness of the weak solution to the system for a large class of initial data on the whole space R 2 .
We first state the assumptions on the initial data and introduce the following notation:
X 0 = { ( n 0 , v 0 , u 0 , c 0 )   n 0 L 2 ( R 2 ) L 1 ( R 2 ) ,   1 + | x | 2 n 0 L 1 ( R 2 ) ,   v 0 H 1 ( R 2 ) , u 0 L 2 ( R 2 ) ,   c 0 L 2 v ( R 2 ) ,   c 0 L 1 ( R 2 ) L ( R 2 ) ,   n 0 0 ,   v 0 0 ,   c 0 0 } .
Now, we state our main theorem.
Theorem 1.
Let ( u 0 ,   v 0 ,   n 0 ,   c 0 ) X 0 and ϕ L ( R 2 ) . Then, the solution of (1) possesses a unique global-in-time weak solution satisfying
n ( L l o c ( R + ; L 2 ( R 2 ) L 1 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) ) ) 2 ,
c L ( R + ; L ( R 2 ) L 1 ( R 2 ) ) L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) ,
u L l o c ( R + ; L 2 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) )
and
v L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) L ( R + ; L 1 ( R 2 ) .
We mention that our results may be generalized to a bounded domain by slightly modifying our proof and adding reasonable boundary conditions. Compared with [5,9], we can obtain the existence and uniqueness of weak solutions to (1) in the whole space R 2 .
Notation. We will set t = t and i = x i for i = 1 , 2 and denote all the partial derivatives β with multi-index β satisfying | β | = k by k   ( k 0 ) . We adopt the convention that the nonessential constant C may change from line to line, and C ( a 1 , a 2 , . . . , a k ) means a constant C depending on a 1 , a 2 , . . . , a k . Given two quantities A and B, we denote B A as B C A . We often label ( a , b ) X = a X + b X .

2. Preliminaries

In the following, we would like to present some preliminaries. We begin with recalling the well-known estimate for the product of two functions.
Lemma 1
([22]). Let s > 0 . Then, there exists a constant C > 0 such that for all u , v L ( R 2 ) H s ( R 2 ) ,
u v H s ( R 2 ) C ( v H s ( R 2 ) u L ( R 2 ) + u H s ( R 2 ) v L ( R 2 ) ) .
A key point of obtaining direct compactness results is the so-called Aubin–Lions lemma.
Lemma 2
(Aubin–Lions [23,24]). Let 1 < α < , 1 β and X 0 , X 1 , X 2 be reflexive separable Banach spaces such that X 0 , X 1 and X 1 X 2 . Then,
{ u L α ( 0 , T ; X 0 ) ; t u L β ( 0 , T ; X 2 ) } L α ( 0 , T ; X 1 ) .
Definition 1.
For T > 0 , let Q T = R 2 × T . For initial data, we say that ( n , m , c , u ) is a weak solution of system (1) if the following conditions hold:
(i)
n ( x , t ) 0 , v ( x , t ) 0 , c ( x , t ) 0 , t 0 , x R 2 , and for any T > 0 ,
n ( L ( [ 0 , T ] ; L 2 ( R 2 ) L 1 ( R 2 ) ) L 2 ( [ 0 , T ] ; H 1 ( R 2 ) ) ) 2 , v ( L ( [ 0 , T ] ; L 2 ( R 2 ) H 1 ( R 2 ) ) L 2 ( [ 0 , T ] ; H 2 ( R 2 ) ) ) 2 , c L ( [ 0 , T ] ; L ( R 2 ) L 1 ( R 2 ) H 1 ( R 2 ) ) L 2 ( [ 0 , T ] ; H 2 ( R 2 ) ) , u L ( [ 0 , T ] ; L 2 ( R 2 ) ) L 2 ( [ 0 , T ] ; H 1 ( R 2 ) ) .
(ii)
Moreover, for any ψ C 0 ( R 2 ; [ 0 , T ] ) ,
Q T n t ψ d x d t + Q T ( u n n ( λ μ n ) ) ψ d x d t = Q T ( χ n c ξ n v n ) ψ d x d t ,
Q T v t ψ d x d t + Q T ( u v + v n ) ψ d x d t = Q T v ψ d x d t ,
Q T c t ψ d x d t + Q T ( u c + n c ) ψ d x d t = Q T c ψ d x d t ,
and for any ψ ˜ C 0 ( R 2 ; [ 0 , T ] ) ,
Q T u t ψ ˜ d x d t + Q T ( u u ( m + n ) ϕ ) ψ ˜ d x d t = Q T u ψ ˜ d x d t
as well as
Q T ψ ˜ · u d x d t = 0 .
If ( v , n , c , u ) is a weak solution of system (1) in R 2 × ( 0 , T ) for any T > 0 , then ( v , n , c , u ) is called a global-in-time weak solution.
Let ( f g ) ( x ) = R 2 f ( x y ) g ( y ) d y . Weak solutions to (1), in the sense of Definition 1, will be constructed as limit objects from a family of appropriately regularized systems as follows:
n t ϵ + ( u ϵ · ) n ϵ = Δ n ϵ χ · ( n ϵ ( c ϵ ρ ϵ ) ) + ξ · ( n ϵ ( v ϵ ρ ϵ ) ) + λ n ϵ μ n ϵ ( n ϵ ρ ϵ ) , c t ϵ + u ϵ · c ϵ = Δ c ϵ ( n ϵ ρ ϵ ) c ϵ , v t ϵ + u ϵ · v ϵ = Δ v ϵ v ϵ + n ϵ ρ ϵ , u t ϵ + ( u ϵ · ) u ϵ = Δ u ϵ + P ϵ ( n ϵ ϕ ) ρ ϵ , · u ϵ = 0 , ( u 0 ϵ , n 0 ϵ , v 0 ϵ , c 0 ϵ ) = ( u 0 ρ ϵ ,   n 0 ρ ϵ ,   v 0 ρ ϵ ,   c 0 ρ ϵ ) ,
where ρ ϵ ( x ) is defined by the standard mollifier ρ ( x ) satisfying R 2 ρ ( x ) d x . We now state the global classical solution for the regularized system to the Cauchy problem (3).
Lemma 3
(Global existence for the regularized system). Let ϕ W ˙ 1 , ( R 2 ) ,
( n 0 ϵ ,   v 0 ϵ ,   c 0 ϵ ,   u 0 ϵ ) ( H s ( R 2 ) ) 4
with s > 1 and n 0 ϵ ,   v 0 ϵ ,   c 0 ϵ 0 . Then, system (3) has a unique global smooth solution
( n ϵ ,   v ϵ ,   c ϵ ,   u ϵ ) ( C ( R + ; H s ( R 2 ) ) ) 4 ( L l o c 2 ( R + ; H s + 1 ( R 2 ) ) ) 4 .
Moreover, n ϵ 0 ,   v ϵ 0 ,   c ϵ 0 for all ( x , t ) R 2 × R + .
Proof. 
The proof of Lemma 3 is standard, we can refer to (Proposition 3.1, [22]). □

3. A Priori Estimates for a Regularized System

In what follows, we let C denote some different constants, which depend at most on χ ,   λ ,   ξ ,   μ ,   ϕ L ( R 2 ) and c 0 L 1 ( R 2 ) L ( R 2 ) ,   v 0 H 1 ( R 2 ) ,   v 0 L 1 ( R 2 ) ,   n 0 L 1 ( R 2 ) ,   u 0 L 2 ( R 2 ) . If there are no special explanations, they are independent of ϵ and t.
Proposition 1.
Assume that ( u 0 , c 0 , m 0 , n 0 ) X 0 , and let ( u ϵ , n ϵ , c ϵ ) be a unique classical solution to the system (3). Then, a positive constant C exists such that
c ϵ ( t ) L 1 ( R 2 ) L ( R 2 ) c 0 L 1 ( R 2 ) L ( R 2 ) , c ϵ ( t ) L 2 ( R 2 ) 2 + 0 t c ϵ ( τ ) L 2 ( R 2 ) 2 d τ c 0 L 2 ( R 2 ) 2
and
n ϵ ( t ) L 1 ( R 2 ) + μ 0 t n ϵ ( t ) L 2 ( R 2 ) 2 d t C n 0 L 1 ( R 2 ) e λ t . , v ϵ ( t ) L 1 ( R 2 ) v 0 L 1 ( R 2 ) + C ( e λ t + t 1 ) n 0 L 1 ( R 2 ) λ + 1
as well as
u ϵ L 2 ( R 2 ) 2 + 0 t u ϵ ( τ ) L 2 ( R 2 ) 2 d τ ( u 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) ) exp ( C t ϕ L ( R 2 ) 2 ) , v ϵ L 2 ( R 2 ) 2 + 0 t v ϵ L 2 ( R 2 ) 2 d τ + 0 t v ϵ L 2 ( R 2 ) 2 d τ v 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) , v ϵ L 2 ( R 2 ) 2 + 0 t v ϵ L 2 ( R 2 ) 2 d τ + 0 t Δ v ϵ L 2 ( R 2 ) 2 d τ ( v 0 L 2 ( R 2 ) 2 + C e t n 0 L 1 ( R 2 ) ) × exp ( u 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) ) 2 exp ( C t ϕ L ( R 2 ) 2 ) .
Proof. 
We first show some priori estimates of n ϵ , v ϵ , and c ϵ . By a direct integrating for, we have
d d t n ϵ ( t ) L 1 ( R 2 ) + μ R 2 n ϵ ( t ) ( n ϵ ( t ) ρ ϵ ) d x = λ n ϵ L 1 ( R 2 ) ,
d d t v ϵ ( t ) L 1 ( R 2 ) + v ϵ ( t ) L 1 ( R 2 ) = n ϵ ρ ϵ L 1 ( R 2 ) ,
c ϵ ( t ) L 1 ( R 2 ) + 0 t ( n ϵ ( τ ) ρ ϵ ) c ϵ ( τ ) L 1 ( R 2 ) d τ = c 0 L 1 ( R 2 ) .
From (4), with the aid of Gronwall’s inequality, we have
n ϵ ( t ) L 1 ( R 2 ) + μ 0 t n ϵ ( t ) L 2 ( R 2 ) 2 d t C n 0 L 1 ( R 2 ) e λ t .
From (5) and (7), we have
v ϵ ( t ) L 1 ( R 2 ) e t v 0 L 1 ( R 2 ) + 0 t e τ t n ρ ϵ L 1 ( R 2 ) d τ e t v 0 L 1 ( R 2 ) + C n 0 L 1 ( R 2 ) 0 t e τ t + λ τ d τ e t v 0 L 1 ( R 2 ) + C n 0 L 1 ( R 2 ) e t λ + 1 ( e λ t + t 1 ) v 0 L 1 ( R 2 ) + C n 0 L 1 ( R 2 ) λ + 1 ( e λ t + t 1 ) .
From (6), the weak maximum principle gives rise to
c ϵ L 1 ( R 2 ) L ( R 2 ) c 0 L 1 ( R 2 ) L ( R 2 ) .
Testing the third equation in (3) by n ϵ and integrating it over R 2 , we obtain
1 2 d d t v ϵ L 2 ( R 2 ) 2 + v ϵ L 2 ( R 2 ) 2 + v ϵ L 2 ( R 2 ) 2 = R 2 ( n ϵ ρ ϵ ) v ϵ d x C n ϵ L 2 ( R 2 ) 2 + 1 2 v ϵ L 2 ( R 2 ) 2 ,
from which, with aid of (8),
v ϵ L 2 ( R 2 ) 2 + 2 0 t v ϵ L 2 ( R 2 ) 2 d τ + 0 t v ϵ L 2 ( R 2 ) 2 d τ v 0 L 2 ( R 2 ) 2 + C 0 t n ϵ L 2 ( R 2 ) 2 d τ v 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) .
Testing the second equation in (3) by c ϵ and integrating it over R 2 , we obtain
1 2 d d t c ϵ L 2 ( R 2 ) 2 + c ϵ L 2 ( R 2 ) 2 = R 2 ( n ϵ ρ ϵ ) ( c ϵ ) 2 d x ,
from which,
c ϵ ( t ) L 2 ( R 2 ) 2 + 2 0 t c ϵ ( τ ) L 2 ( R 2 ) 2 d τ + 2 0 t R 2 n ϵ ( c ϵ ) 2 d x d τ c 0 L 2 ( R 2 ) 2 .
On the other hand, we multiply u ϵ with the fourth equation of (3) and apply the Gagliardo–Nirenberg inequality to deduce that
1 2 d d t u ϵ L 2 ( R 2 ) 2 + u ϵ L 2 ( R 2 ) 2 = R 2 u ϵ · n ϵ ϕ d x u ϵ L 2 ( R 2 ) n ϵ L 2 ( R 2 ) ϕ L ( R 2 ) C u ϵ L 2 ( R 2 ) 2 ϕ L ( R 2 ) 2 + C n ϵ L 2 ( R 2 ) 2 .
From the Gronwall inequality and (7), we have
u ϵ L 2 ( R 2 ) 2 + 0 t u ϵ ( τ ) L 2 ( R 2 ) 2 d τ ( u 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) ) exp ( C t ϕ L ( R 2 ) 2 ) .
It follows from Hölder’s inequality, Gagliardo–Nirenberg’s inequality, and Young’s inequality that
R 2 ( u ϵ · v ϵ ) Δ v ϵ d x u L 4 ( R 2 ) v L 4 ( R 2 ) Δ v L 2 ( R 2 ) C u L 4 ( R 2 ) v L 4 ( R 2 ) 1 / 2 Δ v L 2 ( R 2 ) 3 / 2 C u L 4 ( R 2 ) 4 v L 2 ( R 2 ) 2 + 1 4 Δ v L 2 ( R 2 ) 2 C u L 2 ( R 2 ) 2 u L 2 ( R 2 ) 2 v L 2 ( R 2 ) 2 + 1 4 Δ v L 2 ( R 2 ) 2 .
Applying ∇ to the third equation of (3) gives
t v ϵ + ( u ϵ v ϵ ) Δ v ϵ = v ϵ + ( n ϵ ρ ϵ ) .
Taking the L 2 inner product for above equality with v ϵ and applying (14), we obtain
1 2 d d t v ϵ L 2 ( R 2 ) 2 + v ϵ L 2 ( R 2 ) 2 + Δ v ϵ L 2 ( R 2 ) 2 = R 2 ( u ϵ · v ϵ ) Δ v ϵ d x + R 2 ( n ϵ ρ ϵ ) v ϵ d x C u L 2 ( R 2 ) 2 u L 2 ( R 2 ) 2 v L 2 ( R 2 ) 2 + 1 4 Δ v L 2 ( R 2 ) 2 + 1 4 Δ v L 2 ( R 2 ) 2 + C n L 2 ( R 2 ) 2 ,
from which, from (7), from (13), and from the Gronwall inequality, we have
v ϵ L 2 ( R 2 ) 2 + 0 t v ϵ L 2 ( R 2 ) 2 d τ + 0 t Δ v ϵ L 2 ( R 2 ) 2 d τ ( v 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) ) exp C 0 t u L 2 ( R 2 ) 2 u L 2 ( R 2 ) 2 d t ( v 0 L 2 ( R 2 ) 2 + C e t n 0 L 1 ( R 2 ) ) × exp ( u 0 L 2 ( R 2 ) 2 + C μ 1 e λ t n 0 L 1 ( R 2 ) ) 2 exp ( C t ϕ L ( R 2 ) 2 ) .
Combining this with (7), (8), (9), (10), (11), and (13) directly result. □
Proposition 2.
Suppose v 0 ,   n 0 ,   c 0 0 and that ( u 0 ,   c 0 ,   v 0 ,   n 0 ) X 0 and ϕ L ( R 2 ) . Let ( n ϵ ,   v ϵ ,   c ϵ ,   u ϵ ) be the solutions to the model (3). Then, there a constant C > 0 exists independently of ϵ such that
R 2 n ϵ | ln n ϵ | + | c ϵ | 2 d x + 0 t R 2 | n ϵ | 2 n ϵ d x d τ + 0 t R 2 ( n ϵ ) 2 x d x d τ + 0 t Δ c ϵ L 2 ( R 2 ) 2 + R 2 | c ϵ | 4 ( c ϵ ) 1 d x d τ C e C e C t .
Proof. 
Multiplying Equation (3)2 by ln n ϵ and integrating over R 2 and by prats, we have
d d t R 2 n ϵ ln n ϵ d x + R 2 | n ϵ | 2 n ϵ d x = R 2 n ϵ ( c ϵ ρ ϵ ) ) d x R 2 n ϵ Δ ( v ϵ ρ ϵ ) ) d x + R 2 n ϵ ln n ϵ d x R 2 ( n ϵ ) 2 ln n ϵ d x R 2 n ϵ ( c ϵ ρ ϵ ) ) d x + ϵ n ϵ L 2 ( R 2 ) 2 + C Δ v ϵ L 2 ( R 2 ) 2 + R 2 n ϵ ln n ϵ d x R 2 ( n ϵ ) 2 ln n ϵ d x = R 2 n ϵ ( c ϵ ρ ϵ ) ) d x + ϵ n ϵ L 2 ( R 2 ) 2 + C Δ v ϵ L 2 ( R 2 ) 2 + R 2 n ϵ ln n ϵ d x 0 < n ϵ e x ( n ϵ ) 2 ln n ϵ d x e x < n ϵ < 1 ( n ϵ ) 2 ln n ϵ d x n ϵ 1 ( n ϵ ) 2 ln n ϵ d x = R 2 n ϵ ( c ϵ ρ ϵ ) ) d x + C n ϵ L 2 ( R 2 ) 2 + C Δ v ϵ L 2 ( R 2 ) 2 + R 2 n ϵ ln n ϵ d x 0 < n ϵ e x ( n ϵ ) 2 ln n ϵ d x e x < n ϵ < 1 ( n ϵ ) 2 ln n ϵ d x R 2 n ϵ ( c ϵ ρ ϵ ) ) d x + ϵ n ϵ L 2 ( R 2 ) 2 + C Δ v ϵ L 2 ( R 2 ) 2 + R 2 n ϵ ln n ϵ d x + R 2 ( n ϵ ) 2 x d x + C ,
where x = 1 + x 2 . Since we work in whole space R 2 , here we have to bound R 2 n ϵ ln n ϵ from below. In order to do that, we have to control the behavior of n ϵ as | x | , similarly to [15,25]. To perform this task, we multiply (3)1 by the smooth function x ; then, by integrating it over R 2 , by Young’s inequality, we have
d d t R 2 x n ϵ d x + R 2 ( n ϵ ) 2 x d x = R 2 n ϵ u ϵ x d x + R 2 n ϵ Δ x d x + R 2 n ϵ · c ϵ · x d x ϵ + R 2 n ϵ · v ϵ · x d x ϵ + R 2 n ϵ x d x u ϵ L 2 ( R 2 ) n ϵ L 2 ( R 2 ) x L ( R 2 ) + Δ x L ( R 2 ) n ϵ L 1 ( R 2 ) + R 2 n ϵ x d x + c ϵ L 2 ( R 2 ) n ϵ L 2 ( R 2 ) x L ( R 2 ) + v ϵ L 2 ( R 2 ) n ϵ L 2 ( R 2 ) x L ( R 2 ) C ( u ϵ , c ϵ , v ϵ ) L 2 ( R 2 ) 2 + n ϵ L 1 ( R 2 ) + ϵ n ϵ L 2 ( R 2 ) 2 + R 2 n ϵ x d x .
Multiplying the above inequality by 2 and using (16), one obtains
d d t R 2 ( n ϵ ln n ϵ + 2 x n ) d x + R 2 | n ϵ | 2 n ϵ d x + R 2 ( n ϵ ) 2 x d x R 2 n ϵ ( c ϵ ρ ϵ ) ) d x + C Δ v ϵ L 2 ( R 2 ) 2 + 3 ϵ n ϵ L 2 ( R 2 ) 2 + C ( u ϵ , c ϵ , v ϵ ) L 2 ( R 2 ) 2 + n ϵ L 1 ( R 2 ) + R 2 n ϵ ln n ϵ + 2 n ϵ x d x + C .
By the pointwise identity Δ c ϵ = 2 | c ϵ | 2 + 2 c ϵ Δ c ϵ , the third equation of (3) thereupon turns into the relation
t c ϵ Δ c ϵ ( c ϵ ) 1 | c ϵ | 2 + u ϵ · c ϵ = 1 2 c ϵ ( n ϵ ρ ϵ ) .
Multiplying the above equation by Δ c ϵ , by Young’s inequality, we integrate by parts to obtain
1 2 d d t c ϵ L 2 ( R 2 ) 2 + Δ c ϵ L 2 ( R 2 ) 2 = R 2 u ϵ · c ϵ · Δ c ϵ d x + R 2 1 2 ( n ϵ ρ ϵ ) c ϵ Δ c ϵ x + I 2 = R 2 c ϵ ( c ϵ · ) u ϵ · ( c ϵ ) c ϵ 1 d x + R 2 1 2 ( n ϵ ρ ϵ ) c ϵ Δ c ϵ x + I 2 3 c ϵ L ( R 2 ) u ϵ L 2 ( R 2 ) 2 + 1 12 R 2 | c ϵ | 4 c ϵ d x + R 2 1 2 ( n ϵ ρ ϵ ) c ϵ Δ c ϵ x + I 2 C u ϵ L 2 ( R 2 ) 2 + 1 12 R 2 | c ϵ | 4 c ϵ d x + I 2 R 2 1 4 n ϵ · ( c ϵ ρ ϵ ) + 1 2 n ϵ ρ ϵ | c ϵ | 2 d x ,
where
I 2 = R 2 ( c ϵ ) 1 2 | c ϵ | 2 Δ c ϵ d x .
For I 2 , we have that
I 2 = i , j R 2 ( c ) 1 ( j c ) 2 ( i , i c ) 2 d x = i , j R 2 ( c ϵ ) 2 | i c ϵ | 2 | j c ϵ | 2 d x + 2 i = j R 2 ( c ϵ ) 1 i c ϵ i , j c ϵ j c ϵ d x + 2 i j R 2 ( c ϵ ) 1 i c ϵ i , j c ϵ j c ϵ d x = i , j R 2 ( c ϵ ) 2 | i c ϵ | 2 | j c ϵ | 2 d x 2 I 2 2 i j R 2 ( c ϵ ) 1 ( j c ϵ ) 2 i , i c ϵ d x + 2 i j R 2 ( c ϵ ) 1 i c ϵ i , j c ϵ j c ϵ d x ,
from which, by Young’s inequality, it follows that
I 2 = 1 3 i , j R 2 ( c ϵ ) 2 | i c | 2 | j c ϵ | 2 d x 2 3 i j R 2 ( c ϵ ) 1 ( j c ϵ ) 2 i , i c ϵ d x + 2 3 i j R 2 ( c ϵ ) 1 i c ϵ i , j c ϵ j c ϵ d x 1 3 i , j R 2 ( c ϵ ) 2 | i c ϵ | 2 | j c ϵ | 2 d x + 1 6 i R 2 ( c ϵ ) 2 | i c ϵ | 4 d x + 2 3 i R 2 ( i , i c ϵ ) 2 d x + 1 6 i j R 2 ( c ϵ ) 2 | i c ϵ | 2 | j c | 2 d x + 2 3 i j R 2 ( i , j c ϵ ) 2 d x 1 6 i , j R 2 ( c ϵ ) 2 | i c ϵ | 2 | j c ϵ | 2 d x + 2 3 i , j R 2 ( i , j c ϵ ) 2 d x .
Plugging (21) into (19), we have
1 2 d d t c ϵ L 2 ( R 2 ) 2 + Δ c ϵ L 2 ( R 2 ) 2 + R 2 ( 1 2 n ϵ ρ ϵ ) | c ϵ | 2 d x + 1 6 i , j R 2 ( c ϵ ) 2 | i c | 2 | j c ϵ | 2 d x C u ϵ L 2 ( R 2 ) 2 + 1 12 R 2 | c ϵ | 4 c ϵ d x R 2 1 4 n ϵ · ( c ϵ ρ ϵ ) d x + 2 3 i , j R 2 ( i , j c ϵ ) 2 d x .
Due to Δ c ϵ L 2 ( R 2 ) = 2 c ϵ L 2 ( R 2 ) , we have by multiplying (22) by 4 that
2 d d t c ϵ L 2 ( R 2 ) 2 + 2 R 2 n ϵ ρ ϵ | c | 2 d x + 4 3 Δ c ϵ L 2 ( R 2 ) 2 + 1 3 R 2 | c ϵ | 4 c ϵ d x C u ϵ L 2 ( R 2 ) 2 R 2 n ϵ · ( c ϵ ρ ϵ ) d x .
Summing up (17) and (23), we have
d d t R 2 n ϵ ln n ϵ + 2 x n + 2 | c ϵ | 2 d x + R 2 | n ϵ | 2 n ϵ d x + R 2 ( n ϵ ) 2 x d x + 4 3 Δ c ϵ L 2 ( R 2 ) 2 + 1 3 R 2 | c ϵ | 4 ( c ϵ ) 1 d x + 2 R 2 n ϵ ρ ϵ | c ϵ | 2 d x C n ϵ L 2 ( R 2 ) 2 + u ϵ L 2 ( R 2 ) 2 + Δ v ϵ L 2 ( R 2 ) 2 + ( u ϵ , c ϵ , v ϵ ) L 2 ( R 2 ) 2 + n ϵ L 1 ( R 2 ) + R 2 n ϵ ln n ϵ + 2 x n + 2 | c ϵ | 2 d x ,
from which, let
U 0 = R 2 n 0 ln n 0 + 2 x n 0 + 2 | c 0 | 2 d x ,
we have
R 2 n ϵ ln n ϵ + 2 x n + 2 | c ϵ | 2 d x + 0 t R 2 | n ϵ | 2 n ϵ d x d τ + 0 t R 2 ( n ϵ ) 2 x d x d τ + 0 t Δ c ϵ L 2 ( R 2 ) 2 + R 2 | c ϵ | 4 ( c ϵ ) 1 d x d τ + 0 t R 2 n ϵ ρ ϵ | c ϵ | 2 d x d τ C 0 t ( Δ v ϵ , n ϵ , u ϵ , c ϵ , v ϵ ) L 2 ( R 2 ) 2 + u ϵ L 2 ( R 2 ) 2 + n ϵ L 1 ( R 2 ) d τ + C 0 t R 2 n ϵ ln n ϵ + 2 x n + 2 | c ϵ | 2 d x d τ + C U 0 .
By Proposition 1, one has
0 t ( Δ v ϵ , n ϵ , u ϵ , c ϵ , v ϵ ) L 2 ( R 2 ) 2 + u ϵ L 2 ( R 2 ) 2 + n ϵ L 1 ( R 2 ) d τ C ( e C e C t + e C t + t e C t + 1 ) C e C e C t .
By same reasoning for obtaining (2.27) in [15], we can easily obtain
R 2 n ϵ | ln n ϵ | d x R 2 n ϵ ln n ϵ d x + 4 e 1 R 2 e 1 2 x d x + 2 R 2 n ϵ x d x R 2 n ϵ ln n ϵ d x + C + 2 R 2 n ϵ x d x ,
from (26) and (27) and applying Grönwall’s inequality to (25), we have
R 2 n ϵ | ln n ϵ | + | c ϵ | 2 d x + 0 t R 2 | n ϵ | 2 n ϵ d x d τ + 0 t R 2 ( n ϵ ) 2 x d x d τ + 0 t Δ c ϵ L 2 ( R 2 ) 2 + R 2 | c ϵ | 4 ( c ϵ ) 1 d x d τ C e C t ( C U 0 + e C e C t ) + C C e C e C t .
The proof of Proposition 2 is completed. □
With the preparations of Propositions 1 and 2 at hand, we can further obtain a uniform estimate for the high regularity of ( n ϵ ,   v ϵ ,   c ϵ ,   u ϵ ) .
Proposition 3.
Assume that v 0 ,   n 0 ,   c 0 0 , ( u 0 , c 0 , v 0 , n 0 ) X 0 and ϕ L ( R 2 ) . Let ( n ϵ ,   v ϵ ,   c ϵ ,   u ϵ ) be a solution of system (3). Then, a constant C > 0 exists independently of ϵ such that
c ϵ L t H 1 + c ϵ L t 2 H 2 C e C t
as well as
n ϵ L 2 ( R 2 ) 2 + 0 t n ϵ L 2 ( R 2 ) 2 + n ϵ L 3 ( R 2 ) 3 d s C e C e C e C t .
Proof. 
Now, by the Cauchy–Schwarzed inequality, Propositions 1 and 2, we have
c ϵ L t ( L 2 ( R 2 ) ) + c ϵ L t ( L 2 ( R 2 ) ) = c ϵ L t ( L 2 ( R 2 ) ) + 2 c ϵ c ϵ L t ( L 2 ( R 2 ) ) c ϵ L t ( L 2 ( R 2 ) ) + 2 c ϵ L t ( L ( R 2 ) ) 1 2 c ϵ L t ( L 2 ( R 2 ) ) c 0 L 2 ( R 2 ) + C c 0 L ( R 2 ) 1 2 e C t C e C t .
Using identity Δ c ϵ = 2 c ϵ Δ c ϵ + 2 | c ϵ | 2 again, it follows from Propositions 1 and 2 that
0 t R 2 | Δ c ϵ | 2 d x d τ 8 0 t R 2 c ϵ | Δ c | 2 + | Δ c ϵ | 2 d x d τ 8 c 0 L ( R 2 ) 0 t Δ c ϵ L 2 ( R 2 ) 2 + ( c ϵ ) 1 | c ϵ | 4 L 1 ( R 2 ) d τ C e C t .
Using Proposition 1, we have
0 t c ϵ L 2 ( R 2 ) 2 d τ + 0 t c ϵ L 2 ( R 2 ) 2 d τ t c 0 L 2 ( R 2 ) 2 + 1 2 c 0 L 2 ( R 2 ) 2 .
Then, owing to (29)–(31), we have
c ϵ L t H 1 + c ϵ L t 2 H 2 C e C t .
Testing the first equation in (3) against n ϵ and using Young’s inequality yields
1 2 d d t n ϵ L 2 ( R 2 ) 2 + n ϵ L 2 ( R 2 ) 2 + μ R 2 ( n ϵ ) 2 ( n ϵ ρ ϵ ) d x = λ n ϵ L 2 ( R 2 ) 2 χ 2 R 2 Δ ( c ρ ϵ ) ( n ϵ ) 2 d x + ξ 2 R 2 Δ ( v ρ ϵ ) ( n ϵ ) 2 d x n ϵ L 2 ( R 2 ) 2 + 1 2 ( Δ c ϵ , Δ v ϵ ) L 2 ( R 2 ) n ϵ L 4 ( R 2 ) 2 n ϵ L 2 ( R 2 ) 2 + 1 2 ( Δ c ϵ , Δ v ϵ ) L 2 ( R 2 ) n ϵ L 2 ( R 2 ) n ϵ L 2 ( R 2 ) C 1 + ( Δ c ϵ , Δ v ϵ ) L 2 ( R 2 ) 2 n ϵ L 2 ( R 2 ) 2 + 1 2 n ϵ L 2 ( R 2 ) 2 ,
on the basis of which, (32), Proposition 1, and Grönwall’s inequality, it follows that
n ϵ L 2 ( R 2 ) 2 + 0 t n ϵ L 2 ( R 2 ) 2 + ( n ϵ ) 2 ( n ϵ ρ ϵ ) L 1 ( R 2 ) d s n 0 L 2 ( R 2 ) 2 exp C 0 t 1 + ( Δ c ϵ , Δ v ϵ ) L 2 ( R 2 ) 2 d s n 0 L 2 ( R 2 ) 2 exp C t + C e C t + C e C e C t C e C e C e C t .
Collecting the above inequality with (32), we can thereby complete the proof of Proposition 3. □
Furthermore, using the regularized equations, and the uniform estimates obtained above, we can directly obtain the following proposition.
Proposition 4.
Assume that v 0 ,   n 0   c 0 0 , ( u 0 , c 0 , v 0 , n 0 ) X 0 and ϕ L ( R 2 ) . Let ( n ϵ ,   m ϵ ,   c ϵ ,   u ϵ ) be a solution of system (3). Then,
( t u ϵ , t n ϵ , t c ϵ ) L t 2 ( H 1 ( R 2 ) ) + t v ϵ L t 2 L 2 ( R 2 ) C ( t ) .
Proof. 
We will prove the uniform boundedness for t n ϵ , t v ϵ , t c ϵ and t u ϵ . Making use of (3) and the Gagliardo–Nirenberg inequality
f L t 4 ( L 4 ( R 2 ) ) 2 f L t ( L 2 ( R 2 ) ) f L t 2 ( L 2 ( R 2 ) )
and H s + ϵ ( R 2 ) H s ( R 2 ) for any ϵ > 0 , one can readily obtain that
n t ϵ L t 2 H 1 u ϵ · n ϵ L t 2 ( L 2 ( R 2 ) ) + n ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ ( c ϵ ρ ϵ ) ) L t 2 ( L 2 ( R 2 ) ) + n ϵ ( c ϵ ρ ϵ ) ) L t 2 ( L 2 ( R 2 ) ) + n ϵ ( 1 n ϵ ) L t 2 ( L 2 ( R 2 ) ) u ϵ L t 4 ( L 4 ( R 2 ) ) 2 + n ϵ L t 4 ( L 4 ( R 2 ) ) 2 + n ϵ L t 2 H 1 + c ϵ L t 4 ( L 4 ( R 2 ) ) 2 + v ϵ L t 4 ( L 4 ( R 2 ) ) 2 u ϵ L t ( L 2 ( R 2 ) ) u ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t ( L 2 ( R 2 ) ) n ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 2 ( H 1 ( R 2 ) ) + c ϵ L t ( L 2 ( R 2 ) ) Δ c ϵ L t 2 ( H 2 ( R 2 ) ) + v ϵ L t ( L 2 ( R 2 ) ) Δ v ϵ L t 2 ( H 2 ( R 2 ) ) ,
c t ϵ L t 2 ( H 1 ( R 2 ) ) c ϵ L t 4 ( L 4 ( R 2 ) ) 2 + u ϵ L t 4 ( L 4 ( R 2 ) ) 2 + c ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 4 ( L 4 ( R 2 ) ) 2 c ϵ L t ( L 2 ( R 2 ) ) c ϵ L t 2 ( H 1 ( R 2 ) ) + c ϵ L t 2 ( H 1 ( R 2 ) ) + u ϵ L t L 2 u ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t ( L 2 ( R 2 ) ) n ϵ L t 2 ( H 1 ( R 2 ) )
and
u t ϵ L t 2 ( H 1 ( R 2 ) ) u ϵ L t 4 ( L 4 ( R 2 ) ) 2 + u ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 2 ( L 2 ( R 2 ) ) ϕ L ( R 2 ) u ϵ L t ( L 2 ( R 2 ) ) u ϵ L t 2 ( H 1 ( R 2 ) + u ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 2 ( L 2 ( R 2 ) ) ϕ L ( R 2 )
as well as
v t ϵ L t 2 ( L 2 ( R 2 ) ) v ϵ L t 4 ( L 4 ( R 2 ) ) 2 + u ϵ L t 4 ( L 4 ( R 2 ) ) 2 + v ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 2 ( H 1 ( R 2 ) ) v ϵ L t ( L 2 ( R 2 ) ) v ϵ L t 2 ( H 1 ( R 2 ) ) + v ϵ L t 2 ( H 1 ( R 2 ) ) + n ϵ L t 2 ( H 1 ( R 2 ) ) + u ϵ L t ( L 2 ( R 2 ) u ϵ L t 2 ( H 1 ( R 2 ) ) .
With aid of Proposition 1, Proposition 3 produces the following inequalities:
t u ϵ L t 2 L 2 ( R 2 ) ,   t n ϵ L t 2 L 2 ( R 2 ) ,   t c ϵ L t 2 L 2 ( R 2 ) C ( t )
and t v ϵ L t 2 L 2 ( R 2 ) C ( t ) . This completes the proof of Proposition 4. □

4. Proof of Theorem 1

This section mainly deals with the proof of Theorem 1. In this subsection, we shall extract a suitable subsequence from ( n ϵ ,   v ϵ ,   c ϵ ,   u ϵ ) with the help of a priori energy estimates such that it is convergent, and the corresponding limit triple ( n ,   v ,   c ,   u ) will be a global weak solution of system (1).
  • Existence.
Taking advantage of Propositions 1–4, we can achieve that
n ϵ L t 2 ( H 1 ( R 2 ) ) + + n ϵ L t ( L 2 ( R 2 ) ) + n ϵ L t ( L 1 ( R 2 ) ) C ( t ) ,
c ϵ L 1 ( ( R 2 ) ) L ( ( R 2 ) ) c 0 ϵ L 1 ( R 2 ) L ( R 2 ) , c L t ( H 1 ( R 2 ) ) + c L t 2 ( H 2 ( R 2 ) ) C ( t )
and
u ϵ L t ( L 2 ( R 2 ) ) + u ϵ L t 2 ( H 1 ( R 2 ) ) C ( t )
as well as
v ϵ ( t ) L 1 ( R 2 ) ( n 0 , v 0 ) L 1 ( R 2 ) , v ϵ L t ( H 1 ( R 2 ) ) + v ϵ L t 2 ( H 2 ( R 2 ) ) C ( t ) .
We thus have the following bounds uniformly with ϵ :
n ϵ ( L l o c ( R + ; L 2 ( R 2 ) L 1 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) ) ,
c ϵ L ( R + ; L ( R 2 ) L 1 ( R 2 ) ) L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) ,
u ϵ L l o c ( R + ; L 2 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) )
and
v ϵ L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) L ( R + ; L 1 ( R 2 ) .
The Aubin–Lions compactness Lemma 2 proves that t n ϵ , t u ϵ , t v ϵ are bounded in L l o c 2 ( R + ; H 1 ( R 2 ) ) and t c ϵ is bounded in L l o c 2 ( R + ; L 2 ( R 2 ) ) . Since L 2 ( R 2 ) is locally compactly embedded H s ( R 2 ) and H s ( R 2 ) is continuously embedded in H 1 ( R 2 ) with s ( 1 , 0 ) , by a compactness argument, we thus deduce that a function exists such that for all ϕ ˜ D , the sequence ( ϕ ˜ v ϵ , ϕ ˜ n ϵ , ϕ ˜ c ϵ , ϕ ˜ u ϵ ) converges (up to a subsequence independent of ϕ ˜ ) to ( ϕ ˜ v , ϕ ˜ n , ϕ ˜ c , ϕ ˜ u ) in C ( R + ; H s ( R 2 ) ) . Therefore, in the sense of distribution ( v ϵ , n ϵ , c ϵ , u ϵ ) converges to ( v , n , c , u ) when ϵ 0 . Taking advantage of Fatou’s Lemma, we have
n ( L l o c ( R + ; L 2 ( R 2 ) L 1 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) ) ) 2 ,
c L ( R + ; L ( R 2 ) L 1 ( R 2 ) ) L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) ,
u L l o c ( R + ; L 2 ( R 2 ) ) L l o c 2 ( R + ; H 1 ( R 2 ) )
and
v L l o c ( R + ; H 1 ( R 2 ) ) L l o c 2 ( R + ; H 2 ( R 2 ) ) L ( R + ; L 1 ( R 2 ) .
Hence, we readily obtain that ( v , n , c , u ) is a global-in-time weak solution of system (1).
  • Uniqueness.
Let ( v 1 , n 1 , c 1 , u 1 ) and ( v 2 , n 2 , c 2 , u 2 ) be two solutions of system (1) associated with the same initial data ( v 0 , n 0 , c 0 , u 0 ) . Let V = v 1 v 2 , N = n 1 n 2 , C = c 1 c 2 , and U = u 1 u 2 . From Definition 1, we have for any t > 0 ,
Q t s N ψ d x d s + Q t N · ψ d x d s + Q t ( U · n 1 + u 2 N ) ψ d x d s = ξ Q t ( N v 1 + n 2 V ) ψ d x d s + χ Q t ( N c 1 + n 2 C ) ψ d x d s + Q t ( λ N μ N ( n 1 + n 2 ) ) ψ d x d s .
Taking ψ = N , we have
1 2 N L 2 ( R 2 ) 2 + 0 t N ( s ) L 2 ( R 2 ) 2 d s Q t ( U · n 1 + u 2 N ) N d x d s + Q t ( χ N c 1 + χ n 2 C ξ N v 1 ξ n 2 V ) N d x d s + Q t ( λ N μ N ( n 1 + n 2 ) ) N d x d s = I + I I + I I I .
Using Hölder’s inequality, Young’s inequality, and the Gagliardo–Nirenberg interpolation inequality, by virtue of · u 2 = · U = 0 , we deduce
I = Q t n 1 U · N d x d s C 0 t n 1 L 2 ( R 2 ) n 1 L 2 U L 2 ( R 2 ) U L 2 ( R 2 ) + 1 10 N L 2 ( R 2 ) 2 d s C 0 t n 1 L 2 ( R 2 ) 2 n 1 L 2 ( R 2 ) 2 U L 2 ( R 2 ) 2 d s + 0 t 1 10 U L 2 ( R 2 ) 2 + 1 10 N L 2 ( R 2 ) 2 d s
and
I I C 0 t N L 4 ( R 2 ) 2 c 1 L 4 ( R 2 ) 2 + C n 2 L 4 ( R 2 ) 2 C L 4 ( R 2 ) 2 d s + 0 t C N L 4 ( R 2 ) 2 v 1 L 4 ( R 2 ) 2 + C n 2 L 4 ( R 2 ) 2 V L 4 ( R 2 ) 2 d s + 0 t 1 20 N L 2 ( R 2 ) 2 d s 0 t C N L 2 ( R 2 ) 2 c 1 L 2 ( R 2 ) 2 Δ c 1 L 2 ( R 2 ) 2 + n 2 L 2 ( R 2 ) 2 n 2 L 2 ( R 2 ) 2 C L 2 2 d s + 0 t C N L 2 ( R 2 ) 2 v 1 L 2 ( R 2 ) 2 Δ v 1 L 2 ( R 2 ) 2 + n 2 L 2 ( R 2 ) 2 n 2 L 2 ( R 2 ) 2 V L 2 2 d s + 0 t 1 10 N L 2 ( R 2 ) 2 + 1 10 Δ C L 2 ( R 2 ) 2 + 1 10 Δ V L 2 ( R 2 ) 2 d s
as well as
I I I 0 t C N L 2 ( R 2 ) 2 + C ( n 1 , n 2 ) L 2 ( R 2 ) N L 4 ( R 2 ) 2 d s 0 t C N L 2 ( R 2 ) 2 + C ( n 1 , n 2 ) L 2 N L 2 ( R 2 ) N L 2 d s 0 t C N L 2 ( R 2 ) 2 + C ( n 1 , n 2 ) L 2 ( R 2 ) 2 N L 2 2 d s + 0 t 1 10 N L 2 ( R 2 ) 2 d s .
Substituting (35), (36), and (37) into (34), one has
1 2 N L 2 ( R 2 ) 2 + 0 t N L 2 ( R 2 ) 2 d s 0 t 1 10 ( U , Δ C , Δ V ) L 2 2 + 3 10 N L 2 ( R 2 ) 2 d s + C 0 t ( c 1 L 2 ( R 2 ) 2 Δ c 1 L 2 ( R 2 ) 2 + ( n 1 , n 2 ) L 2 ( R 2 ) 2 + 1 + ( n 1 , n 2 ) L 2 ( R 2 ) 2 ( n 1 , n 2 ) L 2 ( R 2 ) 2 ) ( U , N , C , V ) L 2 ( R 2 ) 2 d s ,
from which we have
N L 2 ( R 2 ) 2 + 0 t N L 2 ( R 2 ) 2 d s 0 t 3 5 ( N , U , Δ C , Δ V ) L 2 2 d s + C 0 t ( c 1 L 2 ( R 2 ) 2 Δ c 1 L 2 ( R 2 ) 2 + ( n 1 , n 2 ) L 2 ( R 2 ) 2 + 1 + ( n 1 , n 2 ) L 2 ( R 2 ) 2 ( n 1 , n 2 ) L 2 ( R 2 ) 2 ) ( U , N , C , V ) L 2 ( R 2 ) 2 d s .
Next from Definition 1, we have
Q t s C ψ d x d s + Q t ( U · c 1 + u 2 C ) ψ d x d s + Q t C · ψ d x d s = Q t ( N c 1 n 2 C ) ψ d x d s .
Taking ψ = C , we have
1 2 C L 2 ( R 2 ) 2 + 0 t C L 2 ( R 2 ) 2 d s 0 t U L 2 ( R 2 ) C L 4 ( R 2 ) c 1 L 4 ( R 2 ) + N L 2 ( R 2 ) c 1 L ( R 2 ) C L 2 ( R 2 ) + n 2 L 2 ( R 2 ) C L 4 ( R 2 ) 2 d s 0 t C U L 2 ( R 2 ) 2 c 1 L 2 ( R 2 ) Δ c 1 L 2 ( R 2 ) + C ( C , N ) L 2 ( R 2 ) 2 d s + 0 t C n 2 L 2 ( R 2 ) 2 C L 2 ( R 2 ) 2 + 1 2 C L 2 ( R 2 ) 2 d s ,
from which we have
C L 2 ( R 2 ) 2 + 0 t C L 2 ( R 2 ) 2 d s C 0 t ( U , N , C ) L 2 ( R 2 ) 2 1 + n 2 L 2 ( R 2 ) 2 + c 1 L 2 ( R 2 ) 2 + Δ c 1 L 2 ( R 2 ) 2 d s .
Since
Q t s i C ψ d x d s + Q t ( i ( U · c 1 ) + u 2 i C + i u 2 C ) ψ d x d s + Q t i C · ψ d x d s = Q t i ( N c 1 + n 2 C ) ψ d x d s .
Letting ψ = i C and summming over i, we infer that
1 2 C L 2 ( R 2 ) 2 + 0 t Δ C L 2 ( R 2 ) 2 d s = Q t ( U · c 1 ) Δ C d x d s Q t ( C · ) u 2 · C d x d s + Q t ( N c 1 + n 2 C ) Δ C d x d s = I I ˜ 1 + I I ˜ 2 + I I ˜ 3 .
Using Hölder’s inequality, Young’s inequality, and the Gagliardo–Nirenberg interpolation inequality, one has
I I ˜ 1 0 t 1 10 Δ C L 2 ( R 2 ) 2 + C U L 4 ( R 2 ) 2 c 1 L 4 ( R 2 ) 2 d s 0 t 1 10 Δ C L 2 ( R 2 ) 2 + 1 10 U L 2 ( R 2 ) 2 + C U L 2 ( R 2 ) 2 c 1 L 2 ( R 2 ) 2 Δ c 1 L 2 ( R 2 ) 2 d s
and
I I ˜ 2 0 t C L 4 ( R 2 ) 2 u 2 L 2 ( R 2 ) d s 0 t 1 10 Δ C L 2 ( R 2 ) 2 + C C L 2 ( R 2 ) 2 u 2 L 2 ( R 2 ) 2 d s
as well as
I I ˜ 3 0 t N L 2 ( R 2 ) Δ C L 2 ( R 2 ) c 1 L ( R 2 ) + n 2 L 4 ( R 2 ) Δ C L 2 ( R 2 ) C L 4 ( R 2 ) d s 0 t C N L 2 ( R 2 ) 2 c 1 L ( R 2 ) 2 + 1 10 Δ C L 2 ( R 2 ) 2 d s + 0 t n 2 L 2 ( R 2 ) 2 n 2 L 2 ( R 2 ) 2 C L 2 ( R 2 ) 2 + 1 10 C L 2 ( R 2 ) 2 d s .
Substituting (43), (44), and (45) into (42), we arrive at
1 2 C L 2 ( R 2 ) 2 + 7 10 0 t Δ C L 2 ( R 2 ) 2 d s C 0 t ( c 1 L 2 ( R 2 ) 2 Δ c 1 L 2 ( R 2 ) 2 + u 2 L 2 ( R 2 ) 2 + 1 + n 2 L 2 ( R 2 ) 2 n 2 L 2 ( R 2 ) 2 ) × ( C , U , N ) L 2 ( R 2 ) 2 d s + 1 10 0 t ( C , U ) L 2 ( R 2 ) 2 d s .
From Definition 1, it follows that
Q t s U ψ ˜ d x d t + Q t ( U · u 1 + u 2 U N ϕ ) ψ ˜ d x d s = Q t U ψ ˜ d x d s
as well as
Q t ψ ˜ · u d x d t = 0 .
Taking ψ ˜ = U yields
1 2 U L 2 ( R 2 ) 2 + 0 t U L 2 ( R 2 ) 2 d s = Q t ( U · U ) · u 1 d x d s + Q t N ϕ · U d x d s C 0 t U L 2 ( R 2 ) 2 ( u 1 L 2 ( R 2 ) 2 u 1 L 2 ( R 2 ) 2 + 1 ) d s + 1 2 0 t U L 2 ( R 2 ) 2 d s + C 0 t N L 2 ( R 2 ) 2 d s ,
which implies that
U L 2 ( R 2 ) 2 + 0 t U L 2 ( R 2 ) 2 d s 0 t C ( U , N ) L 2 ( R 2 ) 2 ( u 1 L 2 ( R 2 ) 2 u 1 L 2 ( R 2 ) 2 + 1 ) d s .
From Definition 1, we have for any t > 0 ,
Q t s V ψ d x d s + Q t ( U · v 1 + u 2 V ) ψ d x d s + Q t V · ψ d x d s = Q t ( N V ) ψ d x d s .
Taking ψ = V , we have
1 2 V L 2 ( R 2 ) 2 + 0 t V L 2 ( R 2 ) 2 d s 0 t C U L 2 ( R 2 ) 2 v 1 L 2 ( R 2 ) Δ v 1 L 2 ( R 2 ) + C V L 2 ( R 2 ) 2 d s + 0 t C u 2 L 2 ( R 2 ) 2 u 2 L 2 ( R 2 ) 2 V L 2 ( R 2 ) 2 + 1 2 V L 2 ( R 2 ) 2 d s ,
from which we have
V L 2 ( R 2 ) 2 + 0 t V L 2 ( R 2 ) 2 d s 0 t C ( U , V ) L 2 ( R 2 ) 2 ( u 2 L 2 ( R 2 ) 2 u 2 L 2 ( R 2 ) 2 + v 1 L 2 ( R 2 ) Δ v 1 L 2 ( R 2 ) + 1 ) d s .
When i = 1 , 2 , we also have
Q t s i V ψ d x d s + Q t ( i ( U · v 1 ) + u 2 i V + i u 2 V ) ψ d x d s + Q t i V · ψ d x d s = Q t i ( N V ) ψ d x d s .
Letting ψ = i V and summming over i, we infer that
1 2 V L 2 ( R 2 ) 2 + 0 t Δ V L 2 ( R 2 ) 2 d s + 0 t V L 2 ( R 2 ) 2 d s = Q t ( U · v 1 ) Δ V d x d s Q t ( V · ) u 2 · V d x d s Q t N Δ V d x d s C 0 t N L 2 ( R 2 ) 2 + U L 4 ( R 2 ) 2 v 1 L 4 ( R 2 ) 2 + u 2 L 2 ( R 2 ) V L 4 ( R 2 ) 2 d s + 1 20 0 t Δ V L 2 ( R 2 ) 2 C 0 t ( N , U , V ) L 2 ( R 2 ) 2 ( 1 + v 1 L 2 ( R 2 ) 2 Δ v 1 L 2 ( R 2 ) 2 + u 2 L 2 ( R 2 ) 2 ) d s + 1 10 0 t Δ V L 2 ( R 2 ) 2 + U L 2 ( R 2 ) 2 d s .
Collecting (39), (42), (46), (50), (42), and (51) leads to
E ˜ ( t ) + 0 t F ˜ ( s ) d s 0 t G ˜ ( s ) E ˜ ( s ) d s ,
with
E ˜ ( t ) = ( U , V , N , C ) ( t ) L 2 ( R 2 ) 2 + ( C , V ) L 2 ( R 2 ) 2 ,
F ˜ ( s ) = ( U , V , N , C ) ( s ) L 2 ( R 2 ) 2 + ( Δ C ( s ) , Δ C ( s ) ) L 2 ( R 2 ) 2 ,
and
G ˜ ( s ) = C ( ( n 1 , n 2 , u 1 , u 2 , c 1 , v 1 ) L 2 2 ( n 1 , n 2 , u 1 , u 2 , Δ c 1 , v 1 ) L 2 2 + ( u 2 , c 1 , Δ c 1 , n 1 , n 2 ) L 2 2 + 1 ) .
Applying Lemmas 1–3, we have
0 t G ˜ ( s ) d s C e C e C t + C e C t + C t ,
from which and (52), applying Grönwall’s inequality, we have
E ˜ ( t ) E ˜ ( t = 0 ) exp 0 t G ˜ ( s ) d s ,
from which we conclude that ( U , V , N , C ) = 0 , and we thus complete the proof of uniqueness.

5. Conclusions

We introduce the notion of a weak solution and establish both the existence and uniqueness of such a weak solution for a large class of initial data on the whole space R 2 .

Author Contributions

Investigation, Y.X. and Q.L.; Supervision, Y.C. and Y.L.; Writing—original draft, M.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This paper is partially supported by the Training Program for Academic and Technical Leaders of Major Disciplines in Jiangxi Province (20204BCJL23057), and the Doctoral Research Fund of Nanchang Normal University (NSBSJJ2023013).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Luca, M.; Chavez-Ross, A.; Edelstein-Keshet, L.; Mogilner, A. Chemotactic signaling, microglia, and Alzheimer’s disease senile plaques: Is there a connection? Bull. Math. Biol. 2003, 65, 693–730. [Google Scholar] [CrossRef] [PubMed]
  2. Painter, K.; Hillen, T. Volume-filling and quorum-sensing in models for chemosensitive movement. Can. Appl. Math. Q. 2002, 10, 501–543. [Google Scholar]
  3. Tuval, I.; Cisneros, L.; Dombrowski, C.; Wolgemuth, C.; Kessler, J.; Goldstein, R. Bacterial swimming and oxygen transport near contact lines. Proc. Natl. Acad. Sci. USA 2005, 102, 2277–2282. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, M. Global solutions to Keller-Segel-Navier–Stokes equations with a class of large initial data in critical Besov spaces. Math. Methods Appl. Sci. 2017, 40, 7425–7437. [Google Scholar] [CrossRef]
  5. Kozono, H.; Miura, M.; Sugiyama, Y. Existence and uniqueness theorem on mild solutions to the Keller-Segel system coupled with the Navier–Stokes fluid. J. Funct. Anal. 2016, 270, 1663–1683. [Google Scholar] [CrossRef]
  6. Liu, J.; Wang, Y. Global weak solutions in a three-dimensional Keller-Segel-Navier–Stokes system involving a tensor-valued sensitivity with saturation. J. Differ. Equ. 2017, 262, 5271–5305. [Google Scholar] [CrossRef]
  7. Tao, Y.; Winkler, M. Global existence and boundedness in a Keller-Segel-Stokes model with arbitrary porous medium diffusion. Discrete Contin. Dyn.Syst. 2012, 32, 1901–1914. [Google Scholar] [CrossRef]
  8. Ren, G.; Liu, B. A new result for global solvability to a two-dimensional attraction-repulsion Navier–Stokes system with consumption of chemoattractant. J. Differ. Equ. 2022, 336, 126–166. [Google Scholar] [CrossRef]
  9. Xie, L.; Xu, Y. Global existence and stabilization in a two-dimensional chemotaxis-Navier–Stokes system with consumption and production of chemosignals. J. Differ. Equ. 2023, 354, 325–372. [Google Scholar] [CrossRef]
  10. Tao, M.Y. Winkler, Locally bounded global solutions in a three-dimensional Chemotaxis-Stokes system with nonlinear diffusion. Ann. Inst. H. Poincaré Anal. Non Linéaire 2013, 30, 157–178. [Google Scholar] [CrossRef]
  11. Cao, X.; Lankeit, J. Global classical small-data solutions for a three-dimensional chemotaxis Navier–Stokes system involving matrix-valued sensitivities. Calc. Var. Partial Differ. 2016, 55, 1339–1401. [Google Scholar] [CrossRef]
  12. Lankeit, J. Long-term behaviour in a chemotaxis-fluid system with logistic source. Math. Models Methods Appl. Sci. 2016, 26, 2071–2109. [Google Scholar] [CrossRef]
  13. Liu, J.; Lorz, A. A coupled chemotaxis-fluid model: Global existence. Ann. Inst. H. Poincaré Anal. Non Linéaire 2011, 28, 643–652. [Google Scholar] [CrossRef]
  14. Lorz, A. Coupled chemotaxis fluid model. Math. Models Methods Appl. Sci. 2010, 20, 987–1004. [Google Scholar] [CrossRef]
  15. Lorz, A. A coupled Keller–Segel–Stokes model: Global existence for small initial data and blow-up delay. Commun. Math. Sci. 2012, 10, 555–574. [Google Scholar] [CrossRef]
  16. Winkler, M. Boundedness and large time behavior in a three-dimensional chemotaxis-Stokes system with nonlinear diffusion and general sensitivity. Comm. Partial Differ. Equ. 2015, 54, 3789–3828. [Google Scholar] [CrossRef]
  17. Winkler, M. Global weak solutions in a three-dimensional Chemotaxis–Navier–Stokes system. Annales l’Institut Henri Poincaré (C) Non Linear Anal. 2015, 10, 555–574. [Google Scholar] [CrossRef]
  18. Winkler, M. Global large-data solutions in a chemotaxis-(Navier-)Stokes system modeling cellular swimming in fluid drops. Commun. Partial. Differ. Equ. 2012, 37, 319–351. [Google Scholar] [CrossRef]
  19. Winkler, M. Stabilization in a two-dimensional chemotaxis-Navier–Stokes system. Arch. Ration. Mech. Anal. 2014, 211, 455–487. [Google Scholar] [CrossRef]
  20. Winkler, M. How far do chemotaxis-driven forces influence regularity in the Navier–Stokes system? Trans. Am. Math. Soc. 2017, 369, 3067–3125. [Google Scholar] [CrossRef]
  21. Yang, M.; Fu, Z.; Sun, J. Existence and large time behavior to coupled chemotaxis-fluid equations in Besov-Morrey spaces. J. Differ. Equ. 2019, 266, 5867–5894. [Google Scholar] [CrossRef]
  22. Zhang, Q.; Zheng, X. Global well-posedness for the two-dimensional incompressible Chemotaxis-Navier–Stokes equations. SIAM J. Math. Anal. 2014, 46, 3078–3105. [Google Scholar] [CrossRef]
  23. Stinner, C.; Tello, J.; Winkler, M. Competitive exclusion in a two-species chemotaxis model. J. Math. Biol. 2014, 68, 1607–1626. [Google Scholar] [CrossRef] [PubMed]
  24. Simon, J. Compact sets in the space Lp(0, T; B). Ann. Mat. Pura Appl. 1987, 146, 65–96. [Google Scholar] [CrossRef]
  25. Duan, R.; Lorz, A.; Markowich, P. Global solutions to the coupled chemotaxis-fuid equations. Comm. Partial Differ. Equ. 2010, 35, 1635–1673. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, Y.; Liu, Q.; Chen, Y.; Lei, Y.; Yang, M. Global Existence of Chemotaxis-Navier–Stokes System with Logistic Source on the Whole Space R2. Axioms 2024, 13, 171. https://doi.org/10.3390/axioms13030171

AMA Style

Xu Y, Liu Q, Chen Y, Lei Y, Yang M. Global Existence of Chemotaxis-Navier–Stokes System with Logistic Source on the Whole Space R2. Axioms. 2024; 13(3):171. https://doi.org/10.3390/axioms13030171

Chicago/Turabian Style

Xu, Yuting, Qianfan Liu, Yao Chen, Yang Lei, and Minghua Yang. 2024. "Global Existence of Chemotaxis-Navier–Stokes System with Logistic Source on the Whole Space R2" Axioms 13, no. 3: 171. https://doi.org/10.3390/axioms13030171

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop