Processing math: 67%
Research article

Existence of Sobolev regular solutions for the incompressible flow of liquid crystals in three dimensions

  • Received: 05 April 2022 Revised: 12 June 2022 Accepted: 21 June 2022 Published: 27 June 2022
  • MSC : 35A01, 35Q31

  • This paper considers a simplified three dimensional Ericksen-Leslie System for nematic liquid crystal flows in the unbounded domain Ω:=R+×R2 or the smooth bounded domain Ω. The hydrodynamic system consists of the Navier-Stokes type equations for the fluid velocity coupled with a convective Ginzburg-Landau type equation for the averaged molecular orientation. We first establish the global existence of Sobolev regular solution with finite energies in Sobolev space Hs(Ω)×Hs(Ω), where the index s of the Sobolev space can be any large fixed integer, but s+. Then we give an asymptotic expansions of a family of Sobolev regularity solutions for such system in Ω.

    Citation: Junling Sun, Xuefeng Han. Existence of Sobolev regular solutions for the incompressible flow of liquid crystals in three dimensions[J]. AIMS Mathematics, 2022, 7(9): 15759-15794. doi: 10.3934/math.2022863

    Related Papers:

    [1] Tariq Mahmood, Zhaoyang Shang . Blow-up criterion for incompressible nematic type liquid crystal equations in three-dimensional space. AIMS Mathematics, 2020, 5(2): 746-765. doi: 10.3934/math.2020051
    [2] Zhongying Liu, Yang Liu, Yiqi Jiang . Global solvability of 3D non-isothermal incompressible nematic liquid crystal flows. AIMS Mathematics, 2022, 7(7): 12536-12565. doi: 10.3934/math.2022695
    [3] Xiufang Zhao . Decay estimates for three-dimensional nematic liquid crystal system. AIMS Mathematics, 2022, 7(9): 16249-16260. doi: 10.3934/math.2022887
    [4] Qiang Li, Mianlu Zou . A regularity criterion via horizontal components of velocity and molecular orientations for the 3D nematic liquid crystal flows. AIMS Mathematics, 2022, 7(5): 9278-9287. doi: 10.3934/math.2022514
    [5] Qiang Li, Baoquan Yuan . Blow-up criterion for the 3D nematic liquid crystal flows via one velocity and vorticity components and molecular orientations. AIMS Mathematics, 2020, 5(1): 619-628. doi: 10.3934/math.2020041
    [6] Xiufang Zhao, Ning Duan . On global well-posedness and decay of 3D Ericksen-Leslie system. AIMS Mathematics, 2021, 6(11): 12660-12679. doi: 10.3934/math.2021730
    [7] Qiang Li, Baoquan Yuan . A regularity criterion for liquid crystal flows in terms of the component of velocity and the horizontal derivative components of orientation field. AIMS Mathematics, 2022, 7(3): 4168-4175. doi: 10.3934/math.2022231
    [8] Zhengyan Luo, Lintao Ma, Yinghui Zhang . Optimal decay rates of higher–order derivatives of solutions for the compressible nematic liquid crystal flows in R3. AIMS Mathematics, 2022, 7(4): 6234-6258. doi: 10.3934/math.2022347
    [9] Yonghui Zou, Xin Xu, An Gao . Local well-posedness to the thermal boundary layer equations in Sobolev space. AIMS Mathematics, 2023, 8(4): 9933-9964. doi: 10.3934/math.2023503
    [10] Danxia Wang, Ni Miao, Jing Liu . A second-order numerical scheme for the Ericksen-Leslie equation. AIMS Mathematics, 2022, 7(9): 15834-15853. doi: 10.3934/math.2022867
  • This paper considers a simplified three dimensional Ericksen-Leslie System for nematic liquid crystal flows in the unbounded domain Ω:=R+×R2 or the smooth bounded domain Ω. The hydrodynamic system consists of the Navier-Stokes type equations for the fluid velocity coupled with a convective Ginzburg-Landau type equation for the averaged molecular orientation. We first establish the global existence of Sobolev regular solution with finite energies in Sobolev space Hs(Ω)×Hs(Ω), where the index s of the Sobolev space can be any large fixed integer, but s+. Then we give an asymptotic expansions of a family of Sobolev regularity solutions for such system in Ω.



    In this paper, we consider the simplified Ericksen-Leslie system, it models the hydrodynamics of nematic liquid crystals in dimension three. It takes the following form:

    tuνΔu+uu+P=κ(dd),u=0,td+ud=γ(df(d)), (1.1)

    where (t,x)R+×Ω with Ω=R+×R2 (or Ω being a bounded smooth domain in R3), u(t,x):R+×ΩR3 is the fluid velocity, P(t,x):R+×ΩR stands for the pressure in the fluid, d(t,x):R+×ΩR3 represents the director field for the averaged macro-scopic molecular orientations. The constants ν, κ and γ are positive constants representing the viscosity of the fluid, the competition between kinetic and potential energy, and the microscopic elastic relaxation time for the molecular orientation field, respectively. The symmetric 3×3 matrix dd denotes the Ericksen stress tensor whose (i,j)-th entry is given by idjd for 1i,j3. The divergence free condition in the second equations of problem (1.1) guarantees the incompressibility of the fluid. The vector valued nonlinear function f(d) is the gradient of certain smooth scalar potential function F(d):R3R such that f(d)=dF(d), we take F to be the Ginzburg-Landau approximation

    F(d)=14ε2(|d|21)2,f(d)=1ε2(|d|21)d,ε>0.

    It is simply the Ginzburg-Landau approximation of the constraint |d| for small ε as given in [1,2].

    The pressure takes the form

    ΔP(t,x)=[uu+κ(dd)]. (1.2)

    Liquid crystals are substances that exhibit a phase of matter that has properties between those of a conventional liquid, and those of a solid crystal. It may flow like a liquid, but its molecules may be oriented in a crystal-like way. In the 1960s, Ericksen [3] and Leslie [4] established the hydrodynamic theory of liquid crystals. The Ericksen-Lislie theory describes the liquid crystal flow, including the velocity vector u and direction vector d of the fluid. Since the general Ericksen-Leslie system is very complicated, a simplified model of the Ericksen-Leslie system [1,2,5] is meanfully to be considered, and there are large diffculties even in the well-posedness theory. One of difficulties is from the three dimension Navier-Stokes type equations in the first equations in (1.1). Another difficulties is from the harmonic maps in the second equations in (1.1) with the constraint condition |d|=1.

    When the direction vector d0, Eq (1.1) is reduced into the famous three dimension incompressible Navier-Stokes equations

    tuνΔu+uu+P=0,u=0. (1.3)

    The question of whether a solution of the 3D incompressible Navier-Stokes equations can develop a finite time singularity from smooth initial data with finite energy is one of the Millennium Prize problems [6]. In 1934, Leray [7] showed that the 3D incompressible Navier-Stokes Eq (1.3) admit global-forward-in-time weak solutions of the initial value problem. Caffarelli et al. [8] established a ε-regularity criterion for Eq (1.3). After that, Lin [9] gave a new and simpler proof for the result of Caffarelli, Kohn and Nirenberg. Buckmaster and Vicol [10] proved that the non-uniqueness of solutions for Eq (1.3) in Hs-space for some small s>0. One can see [11,12,13,14,15,16,17,18] for more results on this equations.

    When the direction vector d0 and f(d)=|d|2d, there is a constraint condition |d|=1. This makes the Ericksen-Leslie system more complicated. For this case, Lin ea al. [19] and Hong [20] established the existence of global weak solutions for this system in R2. Lin and Wang [21] obtained the global existence of a weak solution for the case of three dimension. After that, Huang ea al. [22] showed that two examples of non-trivial solutions to this 3D system with finite time singularity in the unit ball centered at 0.

    In order to relax the constraint condition |d|=1 for the Dirichlet energy

    E(d)=12Ω|d|2dx,

    Lin and Liu [1,2] used the Ginzburg-Landau energy

    E(d)=12Ω(|d|2+12ε2(1|d|2)2)dx

    to replace the Dirichlet energy, then they [1] obtained the global existence of the solution of (1.1) in dimension two and three, after that, they [2] got partial regularity of weak solutions to this system. Furthermore, they [23] proved existence of solutions for system (1.1) and also analyzed the limits of weak solutions of it as ε0. Specially, by setting the term f(d)=0, Hu and Wang [24] proved that all the global weak solutions constructed in [1] be equal to the unique strong solution. Cavaterra et al. [25] established the existence of optimal boundary controls for this system in two dimension. We point out that any solution convergence to the equilibrium state (constants or functions with infinite energy in the unbounded domain) admits the infinite energy. To the author's knowledgement, there is few result on the global finite energy solution with regularity for this system in the unbounded domain or the smooth bounded domain.

    In this paper, we will prove the global existence of Sobolev regular solutions for the three-dimensional incompressible flow of liquid crystals in the unbounded domain and the smooth bounded domain. Meanwhile, we give the explicit asymptotic expansion formula of Sobolev regular solutions for system (1.1).

    We supplement the three-dimensional incompressible nematic liquid crystals (1.1) with the initial data

    u(0,x)=u0(x),d(0,x)=d0(x),xΩ, (1.4)

    and the boundary condition

    u(t,x)=0,d(t,x)=d0(x),(t,x)R+×Ω. (1.5)

    We assume that the initial data (1.4) satisfies the following conditions

    u0(x)=0,xΩ,u0(x)|xΩ=0. (1.6)

    We now state one of main results in this paper.

    Theorem 1.1. Let the parameter ν,γ,κ,ε>0 in (1.1). For any fixed constant s1, if the initial data (1.4) of the three dimension nematic liquid crystal flow (1.1) satisfies the condition (1.6), and there exists a small constant δ(0,1) such that

    u0(x)Hs+2(Ω)δ,d0(x)Hs+2(Ω)δ,

    then the three dimension nematic liquid crystal flow (1.1) with the boundary conditions (1.5) admit a global Sobolev regular solution with finite energy

    (u(t,x),d(t,x))C((0,+);Hs(Ω)×Hs(Ω)).

    Moreover, it holds

    supt(0,+)(u(t,x)Hs(Ω)+d(t,x)Hs(Ω))δ,supt(0,+)P(t,x)Hs(Ω)δ,

    for any (t,x)(0,)×Ω. Here, the pressure is given by (1.2).

    As an application, we have the following asymptotic expansion of solutions.

    Corollary 1.1. Let the parameter ν,γ,κ,ε>0 in (1.1). Assume that the integers p>0,2<q<2(p+1) and the parameter 0<δ1. Assume that the small initial data (1.4) satisfies the condition (1.6). Then the three dimension nematic liquid crystal flow (1.1) in the unbounded domain R+×R2 admit an explicit expansion of global Sobolev regular solution with finite energy as follows

    u(t,x)=(u(0)1(t,x),u(0)2(t,x),u(0)3(t,x))+u0(x)+R1(t,x),d(t,x)=(d(0)1(t,x),d(0)2(t,x),d(0)3(t,x))+d0(x)+R2(t,x),

    for all x=(x1,x2,x3)R+×R2, where

    u(0)1(t,x)=d(0)1(t,x):=δ(1et)xq1x2p+12x2p+13e(x2(p+1)1+x2(p+1)2+x2(p+1)3  ),u(0)2(t,x)=d(0)2(t,x):=δ21(p+1)1(1et)(q2(p+1)x2(p+1)q1)xq11x2p+13e(x2(p+1)1+x2(p+1)2+x2(p+1)3  ),u(0)3(t,x)=d(0)3(t,x):=δ(p+1)1(1et)(q2(p+1)x2(p+1)q1)xq11x2p+12e(x2(p+1)1+x2(p+1)2+x2(p+1)3  ),

    and the remainder term Rk(t,x)Hs(Ω) with s1 and k=1,2 satisfies

    Rk(0,x)=0,R1(t,x)=0,Rk(t,x)|xΩ=0,supt(0,+)Rk(t,x)Hs(Ω)O(δ2).

    Furthermore, one can see that

    u(t,x)|xΩ=0,d(t,x)|xΩ=d0(x),(t,x)R+×Ω.

    Moreover, the pressure is determined by

    ΔP(t,x)=[uu+κ(dd)].

    Remark 1.1. Corollary 1.1 is derived directly from the proof of Theorem 1.1. From Corollary 1.1, we can observe that Eq (1.1) admit a global Sobolev regular solution with finite energy by chosing different constants p and q. Moreover, if the vector functions

    u(0)(t,x)=(u(0)1(t,x),u(0)2(t,x),u(0)3(t,x))

    and

    d(0)(t,x)=(d(0)1(t,x),d(0)2(t,x),d(0)3(t,x))

    satisfy the conditions

    {u(0)(t,x)=0,u(0)(0,x)=0,u(0)Hsδ0,u(0)(t,x)|xΩ=0,

    and

    {d(0)(0,x)=0,d(0)Hsδ0,d(0)(t,x)|xΩ=0.

    Moreover, for any fixed s1 and (t,x)Ω and i,j=1,2,3, it also needs the conditions

    sk=0kxiu(0)j(t,x)Lδ0,

    and

    sk=0kxid(0)j(t,x)Lδ0,

    then we can use the proof of Theorem 1.1 to construct more global Sobolev regular solutions with finite energy of (1.1) in the unbounded domain R+×R2.

    We have the following results.

    Theorem 1.2. Let the parameter ν,γ,κ,ε>0 in (1.1) and γε2<1. For any fixed constant s1, if the initial data (1.4) of the three dimension nematic liquid crystal flow (1.1) satisfies the conditions (1.6), and there exists a small constant 0<δmin{1,ν,γ} such that

    u0(x)Hs+2(Ω)δ,d0(x)Hs+2(Ω)δ,

    then the three dimension nematic liquid crystal flow (1.1) with the boundary conditions (1.5) admit a global Sobolev regular solution with finite energy

    (u(t,x),d(t,x))C((0,+);Hs(Ω)×Hs(Ω)).

    Moreover, it holds

    supt(0,+)(u(t,x)Hs(Ω)+d(t,x)Hs(Ω))δ,supt(0,+)P(t,x)Hs(Ω)δ,

    for any (t,x)(0,)×Ω. Here, the pressure is given by (1.2).

    In particularly, if the domain Ω:=([0,T])3 (a regular polyhedron) with the finite constant T>0, we have the following asymptotic expansion of solutions.

    Corollary 1.2. Let the parameter ν,γ,κ,ε>0 in (1.1) and γε2<1. Assume that the integers p>1 and q>2, and the parameter 0<δ1. The three dimension nematic liquid crystal flow (1.1) in the smooth bounded domain Ω admits an explicit expansion of Sobolev regular solutions with finite energy as follows

    u(t,x)=(u(0)1(t,x),u(0)2(t,x),u(0)3(t,x))+R1(t,x)+u0(x),d(t,x)=(d(0)1(t,x),d(0)2(t,x),d(0)3(t,x))+R1(t,x)+d0(x),

    where x=(x1,x2,x3)Ω:=([0,T])3, and

    u(0)1(t,x)=d(0)1(t,x)=δ(1et)xq1(x1T)qg(x2)g(x3)er(x1,x2,x3),u(0)2(t,x)=d(0)2(t,x)=δ(1et)xq2(x2T)qg(x1)g(x3)er(x1,x2,x3),u(0)3(t,x)=d(0)3(t,x)=2δ(1et)xq3(x3T)qg(x1)g(x2)er(x1,x2,x3),

    with

    g(y):=yq1(yT)q1(2yT)(q2(p+1)y2(p+1)(yT)2(p+1)),r(x1,x2,x3):=3k=1x2(p+1)k(xkT)2(p+1),

    and the remainder term Rk(t,x)Hs(Ω) with s1 and k=1,2 satisfies

    Rk(0,x)=0,R1(t,x)=0,Rk(t,x)|xΩ=0,supt(0,+)Rk(t,x)Hs(Ω)O(δ2).

    Furthermore, one can see that

    u(t,x)|xΩ=0,d(t,x)|xΩ=d0(x),(t,x)R+×Ω.

    Moreover, the pressure is determined by

    ΔP(t,x)=[uu+κ(dd)].

    The organization of this paper is as follows. In Section 2, we show how to choose a suitable initial approximation functions, which lead to the dissipative structure of linearized system. After that, we give the existence of global time-decay Sobolev solution for the linearized equations of the first approximation step. In Section 3, we establish the general approximation step for the construction of the Nash-Moser iteration scheme. This method has been used in [26,27,28,29,30,31]. For the general Nash-Moser implicit function theorem, we refer to the seminal papers of Nash [32], Moser [33] and Hörmander [34]. This last section will show how to construct a global Sobolev solution for the 3D nematic liquid crystal flow (1.1) by using the proof of convergence for the Nash-Moser iteration scheme.

    Throughout this paper, let Ω:=R+×R2 or a smooth bounded domain. We denote the usual norms of L2(Ω) and the Sobolev space Hs(Ω) by L2 and Hs, respectively. The norm of the Sobolev space Hs(Ω):=(Hs(Ω))3 is denoted by Hs. The symbol ab means that there exists a positive constant C such that aCb. We denote by (a,b,c)T the column vector in R3. The letter C with subscripts is used to denote dependencies stands for a positive constant that might change its value at each occurrence.

    The proof of Theorem 1.2 is the same with Theorem 1.1 by a small modification of relationship of parameters ν,γ,ε. Thus we give a proof process including both Theorems 1.1 and 1.2.

    We introduce a family of smooth operators possessing the following properties.

    Lemma 2.1. [34,35] There is a family {Πθ}θ1 of smoothing operators in the space Hs(Ω) acting on the class of functions such that

    ΠθuHs1(Ω)Cθ(s1s2)+uHs2(Ω),s1,s20, (2.1)
    ΠθuuHs1(Ω)Cθs1s2uHs2(Ω),0s1s2,ddθΠθuHs1(Ω)Cθ(s1s2)+1uHs2(Ω),s1,s20,

    where C is a positive constant and (s1s2)+:=max(0,s1s2).

    In our iteration scheme, we set

    θ=Nm=Nm0,m=0,1,2,,

    where N0 is a fixed positive constant, then by (2.1), it holds

    ΠNmuHs1(Ω)Ns1s2muHs2(Ω),s1s2. (2.2)

    We consider the approximation problem of the 3D nematic liquid crystal Eq (1.1) as follows

    L1(u,d):=tuνΔu+ΠNm(uu+P+κ(dd)),L2(u,d):=tdγd+ΠNm(udγε2(|d|21)d), (2.3)

    with the initial data (1.4), the boundary condition (1.5) and the incompressible condition

    u=0.

    Let s1 be a fixed finite constant and 0<δ0<δ21. For any xΩ, we choose the initial approximation functions

    u(0)(t,x)=(u(0)1(t,x),u(0)2(t,x),u(0)3(t,x))Hs(Ω),d(0)(t,x)=(d(0)1(t,x),d(0)2(t,x),d(0)3(t,x))Hs(Ω),

    where we require

    {u(0)(t,x)=0,u(0)(0,x)=0,u(0)Hsδ0,u(0)(t,x)|xΩ=0, (2.4)

    and

    {d(0)(0,x)=0,d(0)Hsδ0,d(0)(t,x)|xΩ=0. (2.5)

    Moreover, for any fixed s1, and t>0 and xΩ and i,j=1,2,3, it also needs the conditions

    sk=0kxiu(0)j(t,x)Lδ0, (2.6)

    and

    sk=0kxid(0)j(t,x)Lδ0, (2.7)

    and the initial error term

    E(0)Hsδ0,¯E(0)Hsδ0, (2.8)

    where E(0) and ¯E(0) denote the error term taking the form

    E(0):=L1(u(0),d(0)),¯E(0):=L2(u(0),d(0)),

    with the vector form

    E(0)=(E(0)1,E(0)2,E(0)3),¯E(0)=(¯E(0)1,¯E(0)2,¯E(0)3).

    There are two family of explicit examples. In fact, many vector functions can be chosen to satisfy conditions (2.4)–(2.8). We now give two families of exact examples of the initial approximation function satisfying (2.4)–(2.8) for the unbounded domain and the smooth bounded domain.

    For the unbounded domain Ω=R+×R2, let the integers p>1 and 2<q<2(p+1). We choose the initial approximation functions of the following form

    u(0)(t,x)=(u(0)1(t,x),u(0)2(t,x),u(0)3(t,x)),(t,x)R+×Ω,d(0)(t,x)=(d(0)1(t,x),d(0)2(t,x),d(0)3(t,x)),(t,x)R+×Ω,

    where

    u(0)1(t,x)=d(0)1(t,x):=δ(1et)xq1x2p+12x2p+13e(x2(p+1)1+x2(p+1)2+x2(p+1)3   ),u(0)2(t,x)=d(0)2(t,x):=δ(1et)(q2(p+1)x2(p+1)q1)xq11x2p+13e(x2(p+1)1+x2(p+1)2+x2(p+1)3   ),u(0)3(t,x)=d(0)3(t,x):=2δ(1et)(q2(p+1)x2(p+1)q1)xq11x2p+12e(x2(p+1)1+x2(p+1)2+x2(p+1)3   ).

    For the smooth bounded domain Ω:=([0,T])3 (a regular polyhedron) with the finite constant T>0, we set the integers p>1 and q>2, and

    u(0)1(t,x)=d(0)1(t,x)=δ(1et)xq1(x1T)qg(x2)g(x3)er(x1,x2,x3),u(0)2(t,x)=d(0)2(t,x)=δ(1et)xq2(x2T)qg(x1)g(x3)er(x1,x2,x3),u(0)3(t,x)=d(0)3(t,x)=2δ(1et)xq3(x3T)qg(x1)g(x2)er(x1,x2,x3),

    with

    g(y):=yq1(yT)q1(2yT)(q2(p+1)y2(p+1)(yT)2(p+1)),r(x1,x2,x3):=3k=1x2(p+1)k(xkT)2(p+1).

    By straightforward computations, we obtain

    u(0)(t,x)=0,

    and

    u(0)(t,x)|xΩ=0,t0,d(0)(t,x)|xΩ=0.

    Moreover, for every fixed integers p and q, the assumptions (2.4)–(2.8) holds.

    Here, the initial approximation pressure satisfies

    ΔP(0)(t,x)=[u(0)u(0)+κ(d(0)d(0))].

    We now construct the first approximation solution denoted by (u(1)(t,λx),d(1)(t,λx)) of system (2.3). There should be two cases on the parameter λ according to the domain.

    Case 2.1. If the domain is the unbounded domain R+×R2, then we set the parameter λ>1.

    Case 2.2. If the domain is a bounded domain Ω, we set the parameter λ=1, meanwhile, we should require λε2<1.

    The first approximation step between the initial approximation function and first approximation solution is denoted by

    h(1)(t,λx):=u(1)(t,λx)u(0)(t,x),w(1)(t,λx):=d(1)(t,λx)d(0)(t,x).

    Then we linearize the nonlinear system (2.3) around (u(0),d(0)) to get the linearized operators as follows

    J1[u0,d(0)](h(1),w(1)):=h(1)tνλ2Δh(1)+ΠN1[λ(u(0))h(1)+(h(1))u(0)+λ((Du(0)P)h(1)+(Dd(0)P)w(1))+κ(d(0)λw(1))+κ(λw(1)d(0))], (2.9)

    and

    J2[u0,d(0)](h(1),w(1)):=w(1)tγλ2Δw(1)+ΠN1[λ(u(0))w(1)+(w(1))u(0)+γε2(|d(0)|2w(1)w(1))], (2.10)

    where Du(0) denotes the Fréchet derivatives on u(0), and by (1.2), it takes the form

    (Du(0)P)h(1):=λΔ13k1=13k2=1(xk2h(1)k1xk1u(0)k2+xk2u(0)k1xk1h(1)k2),(Dd(0)P)w(1):=κλΔ1((λw(1)d(0))+(d(0)λw(1)))=κλΔ1[3k1=13k2=1(2xk1d(0)k2xjw(1)k2+λxk1d(0)k2xk1xjw(1)k2+2xk1w(1)k2xjd(0)k2+λxk1w(1)k2xk1xjd(0)k2)]. (2.11)

    We now consider the linear system

    J1[u0,d(0)](h(1),w(1))=ΠN1E(0),J2[u0,d(0)](h(1),w(1))=ΠN1¯E(0),h(1)=0,h(1)(0,λx)=h(1)0(λx),w(1)(0,λx)=w(1)0(λx), (2.12)

    and the boundary condition

    h(1)(t,λx)|xΩ=0,w(1)(t,λx)|xΩ=0, (2.13)

    from which, the solution of it gives the first approximation step of 3D nematic liquid crystal flow (1.1).

    Before we carry out some a priori estimates, for j=1,2,3, we rewrite equations of (2.12) into a coupled system as follows

    th(1)jνλ2Δh(1)j+λΠN13i=1u(0)ixih(1)j+ΠN13i=1h(1)ixiu(0)j+λΠN1xj((Du(0)P)h(1)+(Dd(0)P)w(1))+κλΠN13k1=13k2=1(2xk1d(0)k2xjw(1)k2+λxk1d(0)k2xk1xjw(1)k2+2xk1w(1)k2xjd(0)k2+λxk1w(1)k2xk1xjd(0)k2)=ΠN1E(0)j, (2.14)

    coupled with

    tw(1)jγλ2Δw(1)j+λΠN13i=1u(0)ixiw(1)j+ΠN13i=1h(1)ixid(0)j+γε2(|d(0)|2w(1)jw(1)j)=ΠN1¯E(0)j, (2.15)

    with the initial data

    h(1)j(0,λx)=h(1)0j(λx),w(1)j(0,λx)=w(1)0j(λx), (2.16)

    and the boundary condition (2.13).

    We will use the weighted function estimate to get the decay in time of solutions. It is like the Carleman estimate. Let ψ(x1) be a function defined in (0,+) such that

    0<κψ(x1)(ψ(x1))2<+, (2.17)

    and eψ(x1) is bounded in (0,+). The condition (2.17) implies ψ(x1)κ>1. In fact, there are many functions can satisfy above conditions. For a simple example, we take the function as the form

    ψ(x1)=ln|cos(κx1)|,x12iπ+π2,foriZ.

    We now derive L2-estimate of solution for the linear systems (2.14) and (2.15).

    Lemma 2.2. If the domain is the unbounded domain R+×R2, let the parameter λ>1. If the domain is a bounded domain Ω, let the parameter λ=1 and γε2<1. Assume that the initial approximation function (u(0),d(0)) satisfies conditions (2.4)–(2.8). Then the solution (h(1)(t,λx),w(1)(t,λx)) of the linear systems (2.14) and (2.15) satisfies

    3j=1Ω((h(1)j)2+(w(1)j)2)dxeCν,γ,λ,ε,δt3j=1(Ω((h(1)0j)2+(w(1)0j)2)dx+ΠN1t0Ω((E(0)j)2+(¯E(0)j)2)dxdt), (2.18)

    where Cν,γ,λ,ε,δ denotes a positive constant depending on ν,γ,λ,ε,δ.

    Proof. Multiplying both sides of equations in (2.14) and (2.15) by eψ(x1)h(1)j and eψ(x1)w(1)j, respectively, then integrating over Ω by noticing the boundary condition (2.13), for j=1,2,3, it follows that

    12ddtΩ(h(1)j)2eψ(x1)dx+νλ23i=1Ω(xih(1)j)2eψ(x1)dx+νλ22Ω(ψ(x1)(ψ(x1))2)(h(1)j)2eψ(x1)dx+λΠN13i=1Ω(u(0)ixih(1)j)h(1)jeψ(x1)dx+ΠN13i=1Ω(h(1)ixiu(0)j)h(1)jeψ(x1)dx+λΠN1Ωxj((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)jeψ(x1)dx+κλΠN13k1=13k2=1Ω((2xk1d(0)k2)(xjw(1)k2)+λ(xk1d(0)k2)(xk1xjw(1)k2)+(2xk1w(1)k2)(xjd(0)k2)+λ(xk1w(1)k2)(xk1xjd(0)k2))h(1)jeψ(x1)dx=ΠN1ΩE(0)jh(1)jeψ(x1)dx, (2.19)

    and

    12ddtΩ(w(1)j)2eψ(x1)dx+γλ23i=1Ω(xiw(1)j)2eψ(x1)dx+γλ22Ω(ψ(x1)(ψ(x1))2)(w(1)j)2eψ(x1)dx+λΠN13i=1Ω(u(0)ixiw(1)j)w(1)jeψ(x1)dx+ΠN13i=1Ω(h(1)ixid(0)j)w(1)jeψ(x1)dx+γε2ΠN1Ω(|d(0)|21)(w(1)j)2eψ(x1)dx=ΠN1Ω¯E(0)jw(1)jeψ(x1)dx. (2.20)

    We sum up (2.19) and (2.20) from j=1 to j=3, hence

    12ddt3j=1Ω((h(1)j)2+(w(1)j)2)eψ(x1)dx+λ23j=13i=1Ω(ν(xih(1)j)2+γ(xiw(1)j)2)eψ(x1)dx+λ223j=1Ω(ψ(x1)(ψ(x1))2)(ν(h(1)j)2+γ(w(1)j)2)eψ(x1)dx+γε2ΠN1Ω(|d(0)|21)(w(1)j)2eψ(x1)dx+λΠN13j=13i=1Ω(u(0)ixih(1)j)h(1)jeψ(x1)dx+ΠN13j=13i=1Ω(h(1)ixiu(0)j)h(1)jeψ(x1)dx+λΠN13j=1Ωxj((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)jeψ(x1)dx+κλΠN13j=13k1=13k2=1Ω((2xk1d(0)k2)(xjw(1)k2)+λ(xk1d(0)k2)(xk1xjw(1)k2) (2.21)
    +(2xk1w(1)k2)(xjd(0)k2)+λ(xk1w(1)k2)(xk1xjd(0)k2))h(1)jeψ(x1)dx+λΠN13j=13i=1Ω(u(0)ixiw(1)j)w(1)jeψ(x1)dx+ΠN13j=13i=1Ω(h(1)ixid(0)j)w(1)jeψ(x1)dx=ΠN13j=1Ω(E(0)jh(1)j+¯E(0)jw(1)j)eψ(x1)dx. (2.22)

    On the one hand, note that we have chosen the initial approximation function (u(0),d(0)) satisfying (2.4)–(2.8). We integrate by parts and we get

    3j=13i=1Ω(u(0)ixih(1)j)h(1)jeψ(x1)dx=123j=13i=1Ωxiu(0)i(h(1)j)2eψ(x1)dx+123j=1Ωψ(x1)u(0)1(h(1)j)2eψ(x1)dx. (2.23)

    Since the initial approximation function u(0) satisfies u(0), inequality (2.23) reduces to

    3j=13i=1Ω(u(0)ixih(1)j)h(1)jeψ(x1)dx=123j=1Ωψ(x1)u(0)1(h(1)j)2eψ(x1)dx. (2.24)

    By direct computation we find

    3j=13i=1Ωh(1)ixiu(0)jh(1)jeψ(x1)dx=3j=1Ωxju(0)j(h(1)j)2eψ(x1)dx+3j=1ijΩh(1)ixiu(0)jh(1)jeψ(x1)dx. (2.25)

    Next, noticing the incompressible condition

    h(1)=0,

    we obtain

    3j=1Ωxj((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)jeψ(x1)dx=3j=1Ω((Du(0)P)h(1)+(Dd(0)P)w(1))xjh(1)jeψ(x1)dx+Ωψ(x1)((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)1eψ(x1)dx=Ωψ(x1)((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)1eψ(x1)dx, (2.26)

    then, we integrate by part to derive

    3k1=13k2=1Ωψ(x1)1(λxk1d(0)k2xk1x1w(1)k2+2xk1w(1)k2x1d(0)k2)h(1)1eψ(x1)dx=23k2=1Ω(ψ(x1)1(x1d(0)k2)(x1w(1)k2)+ψ(x1)1(x1d(0)k2)(x1w(1)k2))h(1)1eψ(x1)dx 3k1=13k2=1Ωψ(x1)1((2xk1d(0)k2)(x1w(1)k2)h(1)1+(xk1d(0)k2)(x1w(1)k2)(xk1h(1)1))eψ(x1)dx 3k1=13k2=1Ωψ(x1)1((xk1x1d(0)k2)(xk1w(1)k2)h(1)1+(x1d(0)k2)(xk1w(1)k2)(xk1h(1)1))eψ(x1)dx, (2.27)

    furthermore, from (2.11), using the standard Calderon-Zygmund theory and Young's inequality, it holds

    |3k1=13k2=1Ωψ(x1)1(λxk1d(0)k2xk1x1w(1)k2+2xk1w(1)k2x1d(0)k2)h(1)1eψ(x1)dx|123k2=1Ω(|ψ(x1)||x1d(0)k2|+|ψ(x1)||x1d(0)k2|+|ψ(x1)|3k1=1(|2xk1d(0)k2|+|xk1x1d(0)k2|))(h(1)1)2eψ(x1)dx+123k1=13k2=1Ω|ψ(x1)|(|xk1d(0)k2|+|x1d(0)k2|)(xk1h(1)1)2eψ(x1)dx+123k1=13k2=1Ω[|ψ(x1)|(|x1d(0)k2|+|2xk1d(0)k2|+|xk1d(0)k2|)+|ψ(x1)||x1d(0)k2|](x1w(1)k2)2eψ(x1)dx+123k1=13k2=1Ω|ψ(x1)|(|xk1x1d(0)k2|+|x1d(0)k2|)(xk1w(1)k2)2eψ(x1)dx, (2.28)

    thus, it holds

    |Ωψ(x1)((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)1eψ(x1)dx|λ|3k1=13k2=1Ωψ(x1)1(xk2h(1)k1xk1u(0)k2+xk2u(0)k1xk1h(1)k2)h(1)1eψ(x1)dx|+κλ|3k1=13k2=1Ωψ(x1)1(2xk1d(0)k2x1w(1)k2+λxk1w(1)k2xk1x1d(0)k2)h(1)1eψ(x1)dx|+κλ|3k1=13k2=1Ωψ(x1)1(λxk1d(0)k2xk1x1w(1)k2+2xk1w(1)k2x1d(0)k2)h(1)1eψ(x1)dx|123k1=13k2=1Ωak1k2(t,x)(h(1)1)2eψ(x1)dx+123k1=13k2=1Ωbk1k2(t,x)(xk1h(1)1)2eψ(x1)dx+123k1=13k2=1Ω|ψ(x1)|(|xk1u(0)k2|(xk2h(1)k1)2+|xk2u(0)k1|(xk1h(1)k2)2)eψ(x1)dx,+123k1=13k2=1Ωck1k2(t,x)(x1w(1)k2)2eψ(x1)dx+123k1=13k2=1Ωek1k2(t,x)(xk1w(1)k2)2eψ(x1)dx, (2.29)

    where

    ak1k2(t,x):=|ψ(x1)|(|xk1u(0)k2|+|xk2u(0)k1|+|2xk1d(0)k2|+|xk1x1d(0)k2|+|x1d(0)k2|)+|ψ(x1)||x1d(0)k2|,bk1k2(t,x):=|ψ(x1)|(|xk1d(0)k2|+|x1d(0)k2|),ck1k2(t,x):=|ψ(x1)|(|x1d(0)k2|+|2xk1d(0)k2|+|xk1d(0)k2|)+|ψ(x1)||x1d(0)k2|,ek1k2(t,x):=|ψ(x1)|(|xk1x1d(0)k2|+|x1d(0)k2|).

    Note that ψ(x1) is bounded in Ω. It follows from inquality (2.27) that

    |Ωψ(x1)((Du(0)P)h(1)+(Dd(0)P)w(1))h(1)1eψ(x1)dx|δΩ(h(1)1)2eψ(x1)dx+δ3j=13i=1Ω(xjh(1)i)2eψ(x1)dx+δ3j=13i=1Ω(xjw(1)xi)2eψ(x1)dx. (2.30)

    On the other hand, we integrate by part to derive

    3j=13k1=13k2=1Ω(λ(xk1d(0)k2)(xk1xjw(1)k2)+(2xk1w(1)k2)(xjd(0)k2))h(1)jeψ(x1)dx=λ3j=13k1=13k2=1Ω((2xk1d(0)k2)(xjw(1)k2)h(1)j+(xk1d(0)k2)(xjw(1)k2)(xk1h(1)j))eψ(x1)dx+λ3j=13k2=1Ωψ(x1)(x1d(0)k2)(xjw(1)k2)hjeψ(x1)dx3j=13k1=13k2=1Ω((xk1w(1)k2)(xk1xjd(0)k2)hj+(xk1w(1)k2)(xjd(0)k2)(xk1hj))eψ(x1)dx+3j=13k2=1Ωψ(x1)(xjd(0)k2)(x1w(1)k2)hjeψ(x1)dx,

    from which, by Young's inequality, we obtain

    κλΠN1|3j=13k1=13k2=1Ω((2xk1d(0)k2)(xjw(1)k2)+λ(xk1d(0)k2)(xk1xjw(1)k2)+(2xk1w(1)k2)(xjd(0)k2)+λ(xk1w(1)k2)(xk1xjd(0)k2))h(1)jeψ(x1)dx|κλδΠN13j=1Ω(h(1)j)2eψ(x1)dx+κλδΠN13j=13k1=1Ω(xk1hj)2eψ(x1)dx+κλδΠN13j=13k2=1Ω(xjw(1)k2)2eψ(x1)dx+κλδΠN13k1=13k2=1Ω(xk1w(1)k2)2eψ(x1)dx, (2.31)

    and

    3j=1ijΩh(1)ixiu(0)jh(1)jeψ(x1)dx12Ω|x2u(0)1+x1u(0)2+x3u(0)1+x1u(0)3|(h(1)1)2eψ(x1)dx+12Ω|x2u(0)1+x1u(0)2+x3u(0)2+x2u(0)3|(h(1)2)2eψ(x1)dx+12Ω|x2u(0)3+x3u(0)2+x3u(0)1+x1u(0)3|(h(1)3)2eψ(x1)dxδ3j=1Ω(h(1)j)2eψ(x1)dx, (2.32)

    and

    λΠN13j=13i=1Ω(u(0)ixiw(1)j)w(1)jeψ(x1)dx+ΠN13j=13i=1Ω(h(1)ixid(0)j)w(1)jeψ(x1)dxλδΠN13j=13i=1Ω((h(1)j)2+(xiw(1)j)2+(w(1)j)2)eψ(x1)dx, (2.33)

    and by ψ(x1)>14, it holds

    3j=1ΩE(0)jh(1)jeψ(x1)dx123j=1Ω((E(0)j)2+|ψ(x1)|(h(1)j)2)eψ(x1)dx,3j=1Ω¯E(0)jw(1)jeψ(x1)dx123j=1Ω((¯E(0)j)2+|ψ(x1)|(w(1)j)2)eψ(x1)dx. (2.34)

    If the domain is the unbounded domain R+×R2, we choose the parameter λ>1. If the domain is a bounded domain Ω, we choose the parameter λ=1 and γε2<1. Next, for a suitable constant λ>1, we substitute (2.23) and (2.34) into (2.21) to get

    ddt3j=1Ω((h(1)j)2+(w(1)j)2)eψ(x1)dx+λ3j=13i=1Ω((νλε)(xih(1)j)2+(γλε)(xiw(1)j)2)eψ(x1)dx+Cλν3j=1Ω(ψ(x1)(ψ(x1))2|ψ(x)|δ)(h(1)j)2eψ(x1)dx+3j=1Ω(γλ(ψ(x1)(ψ(x1))2|ψ(x)|δ)γε2(1δ2))(w(1)j)2eψ(x1)dxΠN13j=1Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dx, (2.35)

    where Cλ denotes a positive constant depending on λ.

    Since ψ(x1) satisfies the ODE inequality (2.17), for suitable positive constants λ and δ, there exist two positive constants Cν,γ,λ and Cν,γ,λ,ε,δ such that

    ψ(x1)(ψ(x1))2|ψ(x)|δ>Cν,γ,λ>0,γλ(ψ(x1)(ψ(x1))2|ψ(x)|δ)γε2(1δ2)>Cν,γ,λ,ε,δ>0,

    thus, integrating (2.35) over (0,t) we find

    3j=1Ω((h(1)j)2+(w(1)j)2)eψ(x1)dx+Cν,γ,λ3j=13i=1t0Ω((xih(1)j)2+(xiw(1)j)2)eψ(x1)dxdt+Cν,γ,λ,ε,δΠN13j=1t0Ω((h(1)j)2+(w(1)j)2)eψ(x1)dxdtΠN13j=1t0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt, (2.36)

    which gives the following inequality

    3j=13i=1t0Ω((xih(1)j)2+(xiw(1)j)2)eψ(x1)dxdtΠN13j=1t0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt.

    Therefore, we apply Gronwall's inequality to inequality (2.36) to obtain

    3j=1Ω((h(1)j)2+(w(1)j)2)eψ(x1)dxeCν,γ,λ,ε,δt3j=1(Ω((h(1)0j)2+(w(1)0j)2)eψ(x1)dx+ΠN1t0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt),

    which yields

    3j=1t0Ω((h(1)j)2+(w(1)j)2)eψ(x1)dxdt3j=1(Ω((h(1)0j)2+(w(1)0j)2)eψ(x1)dx+ΠN1t0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt). (2.37)

    Furthermore, since eψ(x1) is a bounded smooth function in Ω, it follows that

    3j=1Ω((h(1)j)2+(w(1)j)2)dxeCν,γ,λ,ε,δt3j=1(Ω((h(1)0j)2+(w(1)0j)2)dx+ΠN1t0Ω((E(0)j)2+(¯E(0)j)2)dxdt).

    The proof is now complete.

    Next, we derive the higher order derivatives estimates. For a fixed constant s1, we apply Dsi:=sxi (i=1,2,3) to both sides of (2.14). Therefore

    tDsih(1)jνλ2ΔDsih(1)j+λΠN13i=1u(0)ixiDsih(1)j+ΠN13i=1Dsih(1)ixiu(0)j+λΠN1xjDsi((Du(0)P)h(1)+(Dd(0)P)w(1))+κλΠN13k1=13k2=1((2xk1d(0)k2)(Dsixjw(1)k2)+λ(xk1d(0)k2)(Dsixk1xjw(1)k2)+(Dsi2xk1w(1)k2)(xjd(0)k2)+λ(Dsixk1w(1)k2)(xk1xjd(0)k2))=Fj,forj=1,2,3, (2.38)

    coupled with

    tDsiw(1)jγλ2ΔDsiw(1)j+λΠN13i=1u(0)i(Dsixiw(1)j)+ΠN13i=1(Dsih(1)i)xid(0)j+γε2Dsi(|d(0)|2w(1)jw(1)j)=¯Fj,forj=1,2,3, (2.39)

    with the boundary condition

    Dlih(1)j(t,λx)|xΩ=0,Dliw(1)j(t,λx)|xΩ=0, (2.40)

    where the constant 1ls, and

    Fj:=ΠN1DsiE(0)jλΠN1s1+s2=s,0s2s13i=1(Ds1iu(0)i)(xiDs2ih(1)j)ΠN1s1+s2=s,0s2s13i=1(Ds2ih(1)i)(Ds1ixiu(0)j)κλΠN1s1+s2=s,0s2s13k1=13k2=1((Ds1i2xk1d(0)k2)(Ds2ixjw(1)k2)+λ(Ds1ixk1d(0)k2)(Ds2ixk1xjw(1)k2)+(Ds1i2xk1w(1)k2)(Ds2ixjd(0)k2)+λ(Ds1ixk1w(1)k2)(Ds2ixk1xjd(0)k2)),¯Fj:=ΠN1Dsi¯E(0)jλΠN1s1+s2=s,0s2s13i=1(Ds1iu(0)i)(xiDs2iw(1)j)ΠN1s1+s2=s,0s2s13i=1(Ds2ih(1)i)(Ds1ixid(0)j).

    Next, we derive higher derivative estimate of solution for (2.14).

    Lemma 2.3. If the domain is the unbounded domain R+×R2, let the parameter λ>1. If the domain is a bounded domain Ω, let the parameter λ=1 and γε2<1. Assume that the initial approximation function (u(0),d(0)) satisfies conditions (2.4)–(2.8). Then the solution (h(1)(t,λx),w(1)(t,λx)) of the linear system (2.14) and (2.15) satisfies

    3i=13j=1Ω((Dsih(1)j)2+(Dsiw(1)j)2)dxeCμ,γ,λ,ε,δtΠN13j=13i=1sl0=0[Ω((Dl0ih(1)0j)2+(Dl0iw(1)0j)2)dx+0Ω((Dl0iE(0)j)2+(Dl0i¯E(0)j)2)dxdt], (2.41)

    where Cν,γ,λ,ε,δ denotes a positive constant depending on ν,γ,λ,ε,δ.

    Proof. We use an induction argument. Let s=1. By (2.38), we obtain

    tD1ih(1)jνλ2ΔD1ih(1)j+λΠN13i=1u(0)ixiD1ih(1)j+ΠN13i=1D1ih(1)ixiu(0)j+λΠN1xjD1i((Du(0)P)h(1)+(Dd(0)P)w(1))+λΠN13i=1D1iu(0)ixih(1)j+ΠN13i=1h(1)iD1ixiu(0)j+κλΠN13k1=13k2=1D1i((2xk1d(0)k2)(xjw(1)k2)+λ(xk1d(0)k2)(xk1xjw(1)k2)+(2xk1w(1)k2)(xjd(0)k2)+λ(xk1w(1)k2)(xk1xjd(0)k2))=ΠN1D1iE(0)j,forj=1,2,3, (2.42)

    coupled with

    tD1iw(1)jγλ2ΔD1iw(1)j+λΠN13i=1u(0)i(D1ixiw(1)j)+ΠN13i=1(D1ih(1)i)xid(0)j+λΠN13i=1(D1iu(0)i)xiw(1)j+ΠN13i=1h(1)i(D1ixid(0)j)+γε2D1i(|d(0)|2w(1)jw(1)j)=ΠN1D1i¯E(0)j,forj=1,2,3, (2.43)

    with the boundary condition

    D1ih(1)j(t,λx)|xΩ=0,D1iw(1)j(t,λx)|xΩ=0. (2.44)

    We also choose the weighted function satisfies (2.17). Multiplying both sides of (2.42) and (2.43) by D1ih(1)jeψ(x1) and D1iw(1)jeψ(x1), respectively, then integrating over Ω by noticing (2.44), and summing up those equalities from j=1 to j=3, it follows that

    12ddt3j=1Ω(D1ih(1)j)2eψ(x1)dx+νλ23j=13k=1Ω(xkD1ih(1)j)2eψ(x1)dx+νλ223j=1Ω(ψ(x1)(ψ(x1))2)(D1ih(1)j)2eψ(x1)dx+λΠN13j=13i=1Ωu(0)i(xiD1ih(1)j)(D1ih(1)j)eψ(x1)dx+ΠN13j=13i=1ΩD1ih(1)ixiu(0)jD1ih(1)jeψ(x1)dx+λΠN13j=1ΩxjD1i((Du(0)P)h(1)+(Dd(0)P)w(1))D1ih(1)jeψ(x1)dx+ΠN13j=111a=1Ia=0, (2.45)

    and

    12ddt3j=1Ω(D1iw(1)j)2eψ(x1)dx+γλ23j=13k=1Ω(xkD1iw(1)j)2eψ(x1)dx+γλ223j=1Ω(ψ(x1)(ψ(x1))2)(D1iw(1)j)2eψ(x1)dx+λΠN13j=13i=1Ω(u(0)iD1ixiw(1)j)(D1iw(1)j)eψ(x1)dx+ΠN13j=13i=1Ω((D1ih(1)i)xid(0)j)(D1iw(1)j)eψ(x1)dx+λΠN13j=13i=1Ω((D1iu(0)i)xiw(1)j)(D1iw(1)j)eψ(x1)dx+ΠN13j=13i=1Ω(h(1)i(D1ixid(0)j))(D1iw(1)j)eψ(x1)dx+γε2ΠN13j=1ΩD1i(|d(0)|2w(1)jw(1)j)(D1iw(1)j)eψ(x1)dx=ΠN13j=1Ω(D1i¯E(0)j)(D1iw(1)j)eψ(x1)dx, (2.46)

    where

    I1:=λ3j=13i=1ΩD1iu(0)ixih(1)jD1ih(1)jeψ(x1)dx,I2:=3j=13i=1Ωh(1)iD1ixiu(0)jD1ih(1)jeψ(x1)dx,I3:=κλΠN13k1=13k2=1Ω(D1i2xk1d(0)k2)(xjw(1)k2)(D1ih(1)j)eψ(x1)dx,I4:=κλΠN13k1=13k2=1Ω(2xk1d(0)k2)(D1ixjw(1)k2)(D1ih(1)j)eψ(x1)dx,I5:=κλ2ΠN13k1=13k2=1Ω(D1ixk1d(0)k2)(xk1xjw(1)k2)(D1ih(1)j)eψ(x1)dx,I6:=κλ2ΠN13k1=13k2=1Ω(xk1d(0)k2)(D1ixk1xjw(1)k2)(D1ih(1)j)eψ(x1)dx,I7:=κλΠN13k1=13k2=1Ω(D1i2xk1w(1)k2)(xjd(0)k2)(D1ih(1)j)eψ(x1)dx,I8:=κλΠN13k1=13k2=1Ω(2xk1w(1)k2)(D1ixjd(0)k2)(D1ih(1)j)eψ(x1)dx,I9:=κλ2ΠN13k1=13k2=1Ω(D1ixk1w(1)k2)(xk1xjd(0)k2)(D1ih(1)j)eψ(x1)dx,I10:=κλ2ΠN13k1=13k2=1Ω(xk1w(1)k2)(D1ixk1xjd(0)k2)(D1ih(1)j)eψ(x1)dx,I11:=ΠN13j=1Ω((D1iE(0)j)(D1ih(1)j)+(D1i¯E(0)j)(D1iw(1)j))eψ(x1)dx.

    We now estimate each term in (2.45) and (2.46). On the one hand, since we have chosen the initial approximation function (u(0),d(0)) satisfying (2.4)–(2.8), we derive

    3j=13i=1Ωu(0)i(xiD1ih(1)j)(D1ih(1)j)eψ(x1)dx=123j=1Ωψ(x1)u(0)1(D1ih(1)j)2eψ(x1)dx, (2.47)
    3j=13i=1ΩD1ih(1)ixiu(0)jD1ih(1)jeψ(x1)dx=3j=1Ωxju(0)j(D1jh(1)j)2eψ(x1)dx+3j=1ijΩxiu(0)j(D1ih(1)i)(D1ih(1)j)eψ(x1)dx. (2.48)

    By the incompressibility condition h(1)=0 and integration by parts we find

    3j=1ΩxjD1i((Du(0)P)h(1)+(Dd(0)P)w(1))D1ih(1)jeψ(x1)dx=Ωψ(x1)D1i((Du(0)P)h(1)+(Dd(0)P)w(1))(D1ih(1)1)eψ(x1)dx, (2.49)

    from which, similar to (2.28), we use the standard Calderon-Zygmund theory and Young's inequality to derive

    |Ωψ(x1)D1i((Du(0)P)h(1)+(Dd(0)P)w(1))(D1ih(1)1)eψ(x1)dx|δΩ(D1ih(1)1)2eψ(x1)dx+δ3j=13i=1Ω(D1ixjh(1)i)2eψ(x1)dx+δ3j=13i=1Ω(D1ixjw(1)xi)2eψ(x1)dx. (2.50)

    and

    3j=1ijΩxiu(0)j(D1ih(1)i)(D1ih(1)j)eψ(x1)dx12Ω|x2u(0)1+x1u(0)2+x3u(0)1+x1u(0)3|(D1ih(1)1)2eψ(x1)dx+12Ω|x2u(0)1+x1u(0)2+x3u(0)2+x2u(0)3|(D1ih(1)2)2eψ(x1)dx+12Ω|x2u(0)3+x3u(0)2+x3u(0)1+x1u(0)3|(D1ih(1)3)2eψ(x1)dx,

    and

    λΠN13j=13i=1Ω(u(0)iD1ixiw(1)j)(D1iw(1)j)eψ(x1)dxδΠN13j=13i=1Ω((D1ixiw(1)j)2+(D1iw(1)j)2)dx,ΠN13j=13i=1Ω((D1ih(1)i)xid(0)j)(D1iw(1)j)eψ(x1)dxδΠN13j=13i=1Ω((D1ih(1)i)2+(D1iw(1)j)2)dxλΠN13j=13i=1Ω((D1iu(0)i)xiw(1)j)(D1iw(1)j)eψ(x1)dxδΠN13j=13i=1Ω((xiw(1)j)2+(D1iw(1)j)2)dxΠN13j=13i=1Ω(h(1)i(D1ixid(0)j))(D1iw(1)j)eψ(x1)dxδΠN13j=13i=1Ω((h(1)j)2+(D1iw(1)j)2)dxγε2ΠN13j=1ΩD1i(|d(0)|2w(1)jw(1)j)(D1iw(1)j)eψ(x1)γε2(1δ2)ΠN13j=13i=1Ω((D1iw(1)j)2+(w(1)j)2)dx.

    On the other hand, by Young's inequality, it holds

    I1=λ3j=13i=1ΩD1iu(0)i(D1ih(1)j)2eψ(x1)dx,I2δ3j=1Ω((h(1)j)2+(D1ih(1)j)2)eψ(x1)dx,I3δ3k2=1Ω((xjw(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I4δ3k2=1Ω((D1ixjw(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I5δ3k1=13k2=1Ω((xk1xjw(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I8δ3k1=13k2=1Ω((2xk1w(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I9δ3k1=13k2=1Ω((D1ixk1w(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I10δ3k1=13k2=1Ω((xk1w(1)k2)2+(D1ih(1)j)2)eψ(x1)dx,I11123j=1Ω((D1iE(0)j)2+(D1i¯E(0)j)2+|ψ(x1)|(D1ih(1)j)2+|ψ(x1)|(D1iw(1)j)2)eψ(x1)dx, (2.51)

    and we integrate by parts to estimate

    I6δ3k1=13k2=1Ω((xk1xjw(1)k2)2+(D1ih(1)j)2+(D2ih(1)j)2)eψ(x1)dx,I7δ3k1=13k2=1Ω((2xk1w(1)k2)2+(D1ih(1)j)2+(D2ih(1)j)2)eψ(x1)dx. (2.52)

    Thus, summing up (2.45) and (2.46) from i=1 to i=3, we use (2.47)–(2.52) to derive

    ddt3j=13i=1Ω((D1ih(1)j)2+(D1iw(1)j)2)eψ(x1)dx+3j=13i=13k=1Ω((νλδ)(xkD1ih(1)j)2+(γλδ)(xkD1iw(1)j)2)eψ(x1)dx+Cλν3j=13i=1Ω(ψ(x1)(ψ(x1))2|ψ(x)|δ)(D1ih(1)j)2eψ(x1)dx+3j=13i=1Ω(γλ(ψ(x1)(ψ(x1))2|ψ(x)|δ)γε2(1δ2))(D1iw(1)j)2eψ(x1)dxΠN13j=13i=1Ω((D1iE(0)j)2+(D1i¯E(0)j)2)eψ(x1)dx+δΠN13j=1Ω((h(1)j)2+(w(1)j)2)eψ(x1)dx. (2.53)

    If the domain is the unbounded domain R+×R2, we choose the parameter λ>1. If the domain is a bounded domain Ω, we choose the parameter λ=1 and γε2<1. Furthermore, note that ψ(x1) satisfies the ODE inequality (2.17), for suitable positive constants λ and δ, there exist two positive constants Cν,γ,λ and Cν,γ,λ,ε,δ such that we integrate inequality (2.53) over (0,t) to get

    3j=13i=1t0Ω((D1ih(1)j)2+(D1iw(1)j)2)eψ(x1)dxdt+Cν,γ,λ3j=13i=13k=1t0Ω((xkD1ih(1)j)2+(xkD1iw(1)j)2)eψ(x1)dx+Cν,γ,λ,ε,δΠN13j=13i=1t0Ω((D1ih(1)j)2+(D1iw(1)j)2)eψ(x1)dxdtΠN13j=13i=1t0Ω((D1iE(0)j)2+(D1i¯E(0)j)2)eψ(x1)dxdt+δΠN13j=1t0Ω((h(1)j)2+(w(1)j)2)eψ(x1)dxdt. (2.54)

    We observe that the last term in the right-hand side of (2.54) can be controlled by using (2.37). Therefore

    3j=13i=13k=1t0Ω((xkD1ih(1)j)2+(xkD1iw(1)j)2)eψ(x1)dxΠN13j=13i=1t0Ω((D1iE(0)j)2+(D1i¯E(0)j)2)eψ(x1)dxdt+3j=1(Ω((h(1)0j)2+(w(1)0j)2)eψ(x1)dx+ΠN1t0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt). (2.55)

    Hence, by (2.37), we apply Gronwall's inequality to (2.54) to derive

    3j=13i=1t0Ω((D1ih(1)j)2+(D1iw(1)j)2)eψ(x1)dxdteCν,γ,λ,ε,δtΠN13j=1[Ω((h(1)0j)2+(w(1)0j)2)eψ(x1)dx+3i=1Ω((D1ih(1)0j)2+(D1iw(1)0j)2)eψ(x1)dx++0Ω((E(0)j)2+(¯E(0)j)2)eψ(x1)dxdt+3i=10Ω((D1iE(0)j)2+(D1i¯E(0)j)2)eψ(x1)dxdt]. (2.56)

    Assume that the 2\leq l\leq s-1 derivative case holds, that is,

    \begin{equation} \begin{aligned} &\sum\limits_{i = 1}^3\sum\limits_{j = 1}^3\int_{\Omega}\Big((D_i^lh_j^{(1)})^2+(D_i^lw_j^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &\lesssim e^{-C_{\nu,\gamma,\lambda,{\varepsilon},\delta}t}\Pi_{N_1}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{l_0 = 0}^{l}\Big(\int_{\Omega}\Big((D_i^{l_0}h_{0j}^{(1)})^2+(D_i^{l_0}w_{0j}^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &\quad+\int_0^{\infty}\int_{\Omega}\Big((D_i^{l_0}E_j^{(0)})^2+(D_i^{l_0}\overline{E}_j^{(0)})^2\Big)e^{-\psi(x_1)}dxdt\Big), \end{aligned} \end{equation} (2.57)

    and

    \begin{equation} \begin{aligned} &\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{k = 1}^3\int_0^t\int_{\Omega}\Big((\partial _{x_k}D_i^lh_j^{(1)})^2+(\partial _{x_k}D_i^lw_j^{(1)})^2\Big)e^{-\psi(x_1)}dxdt\\ &\lesssim \sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{l_0 = 0}^{l-1}\Big(\int_{\Omega}\Big((D_i^{l_0}h_{0j}^{(1)})^2+(D_i^{l_0}w_{0j}^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &\quad+\Pi_{N_1}\int_0^t\int_{\Omega}\Big((D_i^{l_0}E_j^{(0)})^2+(D_i^{l_0}\overline{E}_j^{(0)})^2\Big)e^{-\psi(x_1)}dxdt\Big). \end{aligned} \end{equation} (2.58)

    We now prove the s th derivative case holds. Multiplying both sides of Eqs (2.38) and (2.39) by D_i^sh^{(1)}_je^{-\psi(x_1)} and D_i^sw^{(1)}_je^{-\psi(x_1)} , then integrating over \Omega by using the boundary condition (2.40), and summing up those equalities from i, j = 1 to i, j = 3 , with similar arguments as for getting (2.53), we can obtain

    \begin{eqnarray} &&{d\over dt}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\int_{\Omega}\Big((D_i^sh_j^{(1)})^2+(D_i^sw_j^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &&+\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\int_{\Omega}\Big((\nu\lambda-\delta)(\partial _{x_j}D_i^sh_j^{(1)})^2+(\gamma\lambda-\delta)(\partial _{x_j}D_i^sw_j^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &&+C_{\nu,\gamma,\lambda,{\varepsilon},\delta}\Pi_{N_1}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\int_{\Omega}\Big((D_i^sh_j^{(1)})^2+(D_i^sw_j^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &\lesssim& \Pi_{N_1}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\int_{\Omega}\Big((D_i^sE_j^{(0)})^2+(D_i^s\overline{E}_j^{(0)})^2\Big)e^{-\psi(x_1)}dx\\ &&+{\varepsilon}\Pi_{N_1}\sum\limits_{k = 0}^{s-1}\sum\limits_{i = 1}^3\int_{\Omega}\Big((D_i^{k}h_i^{(1)})^2+(\partial _{x_i}D_i^{k}h_i^{(1)})^2+(\partial _{x_i}D_i^{k}w_i^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &&+{\varepsilon}\Pi_{N_1}\sum\limits_{k = 0}^{s-1}\sum\limits_{k_1 = 1}^3\sum\limits_{k_2 = 1}^3\Big((D_i^{k}\partial _{x_j}w^{(1)}_{k_2})^2 +(D_i^{k}\partial _{x_{k_1}}^2w^{(1)}_{k_2})^2+(D_i^{k}\partial _{x_{k_1}}w^{(1)}_{k_2})^2\Big),\; \; \; \; \; \; \; \; \end{eqnarray} (2.59)

    where we intergrate by parts to estimate the term (D_i^{s_1}\partial _{x_{k_1}}d^{(0)}_{k_2})(D_i^{s_2}\partial _{x_{k_1}}\partial _{x_j}w^{(1)}_{k_2}) in F_j . Thus, by (2.57) and (2.58), in a similar way as for getting (2.55) and (2.56), applying Gronwall's inequality to (2.59), we obtain

    \begin{equation} \label{YV00-1} \begin{aligned} &\sum\limits_{i = 1}^3\sum\limits_{j = 1}^3\int_{\Omega}\Big((D_i^sh_j^{(1)})^2+(D_i^sw_j^{(1)})^2\Big)e^{-\psi(x_1)}dx\nonumber\\ &\lesssim e^{-C_{\mu,\gamma,\lambda,{\varepsilon},\delta}t}\Pi_{N_1}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{l_0 = 0}^s\Big[\int_{\Omega}\Big((D_i^{l_0}h_{0j}^{(1)})^2+(D_i^{l_0}w_{0j}^{(1)})^2\Big)e^{-\psi(x_1)}dx\\ &\quad+\int_0^{\infty}\int_{\Omega}\Big((D_i^{l_0}E_j^{(0)})^2+(D_i^{l_0}\overline{E}_j^{(0)})^2\Big)e^{-\psi(x_1)}dxdt\Big], \end{aligned} \end{equation}

    and

    \begin{aligned} &\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\int_0^t\int_{\Omega}\Big((\partial _{x_j}D_i^lh_j^{(1)})^2+(\partial _{x_j}D_i^lw_j^{(1)})^2\Big)e^{-\psi(x_1)}dxdt\\ &\lesssim\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{l_0 = 0}^{s-1}\Big(\int_{\Omega}\Big((D_i^{l_0}h_{0j}^{(1)})^2+(D_i^{l_0}w_{0j}^{(1)})^2\Big)^2e^{-\psi(x_1)}dx\\ &\quad+\Pi_{N_1}\int_0^t\int_{\Omega}\Big((D_i^{l_0}E_j^{(0)})^2+(D_i^{l_0}\overline{E}_j^{(0)})^2\Big)e^{-\psi(x_1)}dxdt\Big). \end{aligned}

    Noticing that e^{-\psi(x_1)} is be a bounded function in (0, +\infty) , it follows from inequality (2.60) that

    \begin{aligned} \sum\limits_{i = 1}^3\sum\limits_{j = 1}^3\int_{\Omega}\Big((D_i^sh_j^{(1)})^2+(D_i^sw_j^{(1)})^2\Big)dx&\lesssim e^{-C_{\mu,\gamma,\lambda,{\varepsilon},\delta}t}\Pi_{N_1}\sum\limits_{j = 1}^3\sum\limits_{i = 1}^3\sum\limits_{l_0 = 0}^s\Big[\int_{\Omega}\Big((D_i^{l_0}h_{0j}^{(1)})^2+(D_i^{l_0}w_{0j}^{(1)})^2\Big)dx\\ &\quad+\int_0^{\infty}\int_{\Omega}\Big((D_i^{l_0}E_j^{(0)})^2+(D_i^{l_0}\overline{E}_j^{(0)})^2\Big)dxdt\Big]. \end{aligned}

    The proof is now complete.

    Based on above a priori estimates, we are ready to prove the existence of the first approximation step by using semigroup theory arguments, see [36].

    Proposition 2.1. If the domain is the unbounded domain \mathbb R^+\times \mathbb R^2 , let the parameter \lambda > 1 . If the domain is a bounded domain \Omega , let the parameter \lambda = 1 and {\gamma\over{\varepsilon}^2} < 1 . Assume the initial approximation function (u^{(0)}, d^{(0)}) satisfying (2.4)–(2.8). Then the linearized system

    \begin{equation} \left\{ \begin{array}{lll} & \mathcal J_1[u^{0},d^{(0)}](\boldsymbol{h}^{(1)},\boldsymbol{w}^{(1)}) = \Pi_{N_1}E^{(0)},\\ & \mathcal J_2[u^{0},d^{(0)}](\boldsymbol{h}^{(1)},\boldsymbol{w}^{(1)}) = \Pi_{N_1}\overline{E}^{(0)},\\ &\nabla\cdot \boldsymbol{h}^{(1)} = 0,\\ &\boldsymbol{h}^{(1)}(0,\lambda x) = 0,\quad \boldsymbol{w}^{(1)}(0,\lambda x) = 0, \end{array} \right. \end{equation} (2.60)

    and the boundary condition

    \boldsymbol{h}^{(1)}(t,\lambda x)|_{x\in\partial \Omega} = 0,\quad \boldsymbol{w}^{(1)}(t,\lambda x)|_{x\in\partial \Omega} = 0,

    admit a global solution

    (\boldsymbol{h}^{(1)}(t,\lambda x),\boldsymbol{h}^{(1)}(t,\lambda x))\in \mathbb C((0,+\infty);H^s(\Omega)\times H^s(\Omega)).

    Moreover, the global solution satisfies

    \begin{equation} \|\boldsymbol{h}^{(1)}\|^2_{H^s}+\|\boldsymbol{w}^{(1)}\|^2_{H^s} \lesssim \|\Pi_{N_1}E^{(0)}\|^2_{H^s}+\|\Pi_{N_1}\overline{E}^{(0)}\|^2_{H^s},\quad \forall t > 0. \end{equation} (2.61)

    Proof. The idea of proof is based on the semigroup theory. We first rewrite it as an evolution equation, then by the semigroup generator given in [36], we know the evolution operator generate a semigroup, a local existence can been shown. It combine with the dissipative energy estimate given in Lemma 2.3, the global existence of linear evolution is obtained. More precisely, let \mathbb P be the Leray projector onto the space of divergence free functions. We apply the Leray projector to Eq (2.12). Therefore

    \begin{equation} \begin{aligned} &\textbf{h}^{(1)}_t-\nu\lambda^2 \mathbb P\Delta \textbf{h}^{(1)} + \mathbb N_1(\textbf{h}^{(1)},\textbf{w}^{(1)}) = \mathbb P\Pi_{N_1}E^{(0)},\\ &\textbf{w}^{(1)}_t-\gamma\lambda^2 \mathbb P\Delta\textbf{w}^{(1)}+ \mathbb N_2(\textbf{h}^{(1)},\textbf{w}^{(1)}) = \mathbb P\Pi_{N_1}\overline{E}^{(0)}, \end{aligned} \end{equation} (2.62)

    where

    \begin{aligned} \mathbb N_1(\textbf{h}^{(1)},\textbf{w}^{(1)})&: = \mathbb P\Pi_{N_1}\Big[\lambda\Big(u^{(0)}\cdot\nabla\Big)\textbf{h}^{(1)}+\Big(\textbf{h}^{(1)}\cdot\nabla\Big)u^{(0)}\\ &\quad+\kappa\nabla\cdot(\nabla d^{(0)}\odot\lambda\nabla\textbf{w}^{(1)})+\kappa\nabla\cdot(\lambda\nabla \textbf{w}^{(1)}\odot\nabla d^{(0)})\Big]. \end{aligned}

    and

    \begin{aligned} \mathbb N_2(\textbf{h}^{(1)},\textbf{w}^{(1)})&: = \mathbb P\Pi_{N_1}\Big[\lambda\Big(u^{(0)}\cdot\nabla\Big)\textbf{w}^{(1)}+\Big(\textbf{w}^{(1)}\cdot\nabla\Big)u^{(0)}\\ &\quad+\gamma\Big(|\nabla d^{(0)}|^2\textbf{w}^{(1)}+2\lambda(\nabla d^{(0)},\nabla \textbf{w}^{(1)})d^{(0)}\Big)\Big]. \end{aligned}

    For convenience, we rewrite (2.62) as

    \begin{equation} {d\over dt}\left( \begin{array}{cccc} \textbf{h}^{(1)}\\ \textbf{w}^{(1)} \end{array} \right) + \left( \begin{array}{cccc} -\nu\lambda^2 \mathbb P\triangle&0\\ 0&-\gamma\lambda^2 \mathbb P\triangle \end{array} \right) \left( \begin{array}{cccc} \textbf{h}^{(1)}\\ \textbf{w}^{(1)} \end{array} \right) + \left( \begin{array}{cccc} \mathbb N_1(\textbf{h}^{(1)},\textbf{w}^{(1)})\\ \mathbb N_2(\textbf{h}^{(1)},\textbf{w}^{(1)}) \end{array} \right) = \left( \begin{array}{cccc} \mathbb P\Pi_{N_1}E^{(0)}\\ \mathbb P\Pi_{N_1}\overline{E}^{(0)} \end{array} \right). \end{equation} (2.63)

    We notice that the term \left(\begin{array}{cccc} \mathbb N_1(\textbf{h}^{(1)}, \textbf{w}^{(1)})\\ \mathbb N_2(\textbf{h}^{(1)}, \textbf{w}^{(1)}) \end{array} \right) can be seen as a bounded perturbation of the linear operator ML: = \left(\begin{array}{cccc} -\nu\lambda^2 \mathbb P\triangle & 0\\ 0 & -\gamma\lambda^2 \mathbb P\triangle \end{array} \right) . We follow the idea of [36,37] to show that the linear operator

    \mathbb Z: = ML+BP

    can generate a strongly continuous semigroup e^{ \mathbb Z \tau} in Sobolev space H^s(\Omega)\times H^s(\Omega) by the Lumer-Phillips theorem [38]. Hence linear Eq (2.60) has a solution in \mathbb C((0, +\infty); H^s(\Omega)\times H^s(\Omega)) . Furthermore, it follows from Lemmas 2.2 and 2.3 that the estimate (2.61) holds.

    Let {\varepsilon}\in(0, 1) be a fixed constant. We define

    \begin{equation} \mathcal B_{{\varepsilon}}: = \{(u^{(k)}(t,\lambda x),d^{(k)}(t,\lambda x)): \|u^{(k)}\|_{H^s}+\|d^{(k)}\|_{H^s}\lesssim \delta < 1\} \end{equation} (3.1)

    with the integer 2\leq k\leq m-1 and the constant s\geq1 .

    Assume that the m -th approximation solutions of (2.3) is denoted by (\textbf{h}^{(m)}(t, \lambda x), \textbf{w}^{(m)}(t, \lambda x)) with m = 2, 3, \ldots . Let

    \begin{aligned} \textbf{h}^{(m)}(t,\lambda x)&: = u^{(m)}(t,\lambda x)-u^{(m-1)}(t,\lambda x),\\ \textbf{w}^{(m)}(t,\lambda x)&: = d^{(m)}(t,\lambda x)-d^{(m-1)}(t,\lambda x), \end{aligned}

    it holds

    \begin{aligned} &u^{(m)}(t,\lambda x) = u^{(0)}(t,x)+\textbf{h}^{(1)}(t,\lambda x)+\sum\limits_{i = 2}^m\textbf{h}^{(i)}(t,\lambda x),\\ &d^{(m)}(t,\lambda x) = d^{(0)}(t,x)+\textbf{w}^{(1)}(t,\lambda x)+\sum\limits_{i = 2}^m\textbf{w}^{(i)}(t,\lambda x). \end{aligned}

    We linearize the nonlinear Eq (2.3) around (u^{(m-1)}(t, \lambda x), d^{(m-1)}(t, \lambda x)) to get the following initial value problem

    \begin{equation} \left\{ \begin{array}{lll} & \mathcal J_1[u^{m-1},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)}) = \Pi_{N_m}E^{(m-1)},\\ & \mathcal J_2[u^{m-1},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)}) = \Pi_{N_m}\overline{E}^{(m-1)},\\ &\nabla\cdot \textbf{h}^{(m)} = 0,\\ &\textbf{h}^{(m)}(0,\lambda x) = 0,\quad \textbf{w}^{(m)}(0,\lambda x) = 0, \end{array} \right. \end{equation} (3.2)

    with the boundary conditions

    \begin{equation} \textbf{h}^{(m)}(t,\lambda x)|_{x\in\partial \Omega} = 0,\quad \textbf{w}^{(m)}(t,\lambda x)|_{x\in\partial \Omega} = 0 \end{equation} (3.3)

    where the error term is

    \begin{equation} \begin{aligned} E^{(m-1)}&: = \mathcal L_1[u^{m-1}(t,\lambda x),d^{m-1}(t,\lambda x)] = \mathcal R_1(\textbf{h}^{(m)},\textbf{w}^{(m)}),\\ \overline{E}^{(m-1)}&: = \mathcal L_2[u^{m-1}(t,\lambda x),d^{m-1}(t,\lambda x)] = \mathcal R_2(\textbf{h}^{(m)},\textbf{w}^{(m)}), \end{aligned} \end{equation} (3.4)

    and

    \begin{equation} \begin{aligned} \mathcal R_1(\textbf{h}^{(m)},\textbf{w}^{(m)})&: = \mathcal L_1(u^{(m-1)}+\textbf{h}^{(m)},d^{(m-1)}+\textbf{w}^{(m)})- \mathcal L_1(u^{(m-1)},d^{(m-1)})\\ &\quad- \mathcal J_1[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)})\\ & = \Pi_{N_m}\Big(\lambda\textbf{h}^{(m)}\cdot\nabla\textbf{h}^{(m)}+\nabla P^{(m)}+\kappa\lambda^2\nabla\cdot(\nabla \textbf{w}^{(m)}\odot\nabla \textbf{w}^{(m)})\Big), \end{aligned} \end{equation} (3.5)

    and

    \begin{equation} \begin{aligned} \mathcal R_2(\textbf{h}^{(m)},\textbf{w}^{(m)})&: = \mathcal L_2(u^{(m-1)}+\textbf{h}^{(m)},d^{(m-1)}+\textbf{w}^{(m)})- \mathcal L_2(u^{(m-1)},d^{(m-1)})\\ &\quad- \mathcal J_2[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)})\\ & = \lambda\Pi_{N_m}\Big(\textbf{h}^{(m)}\cdot\nabla\textbf{w}^{(m)}-{\gamma\over{\varepsilon}^2}(|\textbf{w}^{(m)}|^2-1)\textbf{w}^{(m)}\Big), \end{aligned} \end{equation} (3.6)

    This is also the nonlinear term in the approximation problem (2.3) at (u^{(m-1)}(t, \lambda x), d^{(m-1)}(t, \lambda x)) . Here the approximation pressure satisfies

    \Delta P^{(m)}(t,x) = -\nabla\cdot\Big[u^{(m)}\cdot\nabla u^{(m)}+\kappa\nabla\cdot(\nabla d^{(m)}\odot\nabla d^{(m)})\Big].

    The following result describes how to construct the m -th approximation solution.

    Proposition 3.1. If the domain is the unbounded domain \mathbb R^+\times \mathbb R^2 , let the parameter \lambda > 1 . If the domain is a bounded domain \Omega , let the parameter \lambda = 1 and {\gamma\over{\varepsilon}^2} < 1 . Assume the initial approximation function (u^{(0)}, d^{(0)}) satisfying (2.4)–(2.8). (u^{(m-1)}, d^{(m-1)})\in \mathcal B_{{\varepsilon}} and \sum_{i = 1}^{m-1}\Big(\|\textbf{h}^{(i)}\|^2_{H^s}+\|\textbf{w}^{(i)}\|^2_{H^s}\Big)\lesssim\delta^2 . Then the linearized problem (3.2) with the boundary condition (3.3) admits a global solution

    (\boldsymbol{h}^{(m)}(t,\lambda x),\boldsymbol{w}^{(m)}(t,\lambda x))\in \mathbb C((0,+\infty);H^s(\Omega)\times H^s(\Omega)),

    which satisfies

    \begin{equation} \|\boldsymbol{h}^{(m)}\|^2_{H^s}+\|\boldsymbol{w}^{(m)}\|^2_{H^s} \lesssim \|\Pi_{N_m}E^{(m-1)}\|^2_{H^s}+\|\Pi_{N_m}\overline{E}^{(m-1)}\|^2_{H^s},\quad \forall t > 0, \end{equation} (3.7)

    where the error term verifies

    \begin{equation} \begin{aligned} \|E^{(m)}\|_{H^s}+\|\overline{E}^{(m)}\|_{H^s} & = \| \mathcal R_1(\boldsymbol{h}^{(m)},\boldsymbol{w}^{(m)})\|_{H^s}+\| \mathcal R_2(\boldsymbol{h}^{(m)},\boldsymbol{w}^{(m)})\|_{H^s}\\ &\lesssim N_m^3\Big(\|\boldsymbol{h}^{(m)}\|^2_{H^s}+\|\boldsymbol{w}^{(m)}\|^2_{H^s}+\|\boldsymbol{w}^{(m)}\|^4_{H^s}\Big). \end{aligned} \end{equation} (3.8)

    Proof. By direct computation we find

    \begin{equation} \begin{aligned} \partial _{x_i}u^{(m-1)}_j(t,x)& = \partial _{x_i}u^{(0)}_j(t,x)+\lambda\partial _{x_i}\textbf{h}^{(1)}(t,\lambda x)+\lambda\sum\limits_{i = 2}^{m-1}\partial _{x_i}\textbf{h}^{(i)}(t,\lambda x),\\ \partial _{x_i}d^{(m-1)}_j(t,x)& = \partial _{x_i}d^{(0)}_j(t,x)+\lambda\partial _{x_i}\textbf{w}^{(1)}(t,\lambda x)+\lambda\sum\limits_{i = 2}^{m-1}\partial _{x_i}\textbf{w}^{(i)}(t,\lambda x). \end{aligned} \end{equation} (3.9)

    By the assumption

    \sum\limits_{i = 1}^{m-1}\Big(\|\textbf{h}^{(i)}\|^2_{H^s}+\|\textbf{w}^{(i)}\|^2_{H^s}\Big)\lesssim\delta^2,

    for a fixed constant \lambda > 1 , it follows that

    \begin{aligned} &\partial _{x_i}u^{(m-1)}_j(t,x)\sim\partial _{x_i}u^{(0)}_j(t,x)+ \mathcal O(\delta^2),\\ &\partial _{x_i}d^{(m-1)}_j(t,x)\sim\partial _{x_i}d^{(0)}_j(t,x)+ \mathcal O(\delta^2). \end{aligned}

    Thus, noticing that (u^{(0)}(t, x), d^{(0)}(t, x)) satisfies (2.4)–(2.8), by small modification of \partial _{x_i}u^{(0)}_j(t, x) and \partial _{x_i}d^{(0)}_j(t, x)) , it follows that

    \begin{equation} \sum\limits_{k = 0}^s\|\Pi_{N_m}\partial _{x_i}^{k}u^{(m-1)}_j(t,x)\|_{L^{\infty}}\lesssim\delta_0, \end{equation} (3.10)

    and

    \begin{equation} \sum\limits_{k = 0}^s\|\Pi_{N_m}\partial _{x_i}^{k}d^{(m-1)}_j(t,x)\|_{L^{\infty}}\lesssim\delta_0. \end{equation} (3.11)

    Moreover, we notice that the (m-1) -th approximation solution is

    \begin{aligned} &u^{(m-1)}(t,{\varepsilon} x) = u^{(0)}(t,x)+\textbf{h}^{(1)}(t,\lambda x)+\sum\limits_{i = 2}^{m-1}\textbf{h}^{(i)}(t,\lambda x),\\ &d^{(m-1)}(t,{\varepsilon} x) = d^{(0)}(t,x)+\textbf{w}^{(1)}(t,\lambda x)+\sum\limits_{i = 2}^{m-1}\textbf{w}^{(i)}(t,\lambda x), \end{aligned}

    and

    \nabla\cdot \textbf{h}^{(m-1)} = 0.

    We obtain

    \left\{ \begin{array}{lll} &&\nabla\cdot u^{(m-1)}(t,x) = 0,\\ &&u^{(m-1)}(0,x) = 0,\\ &&\|u^{(m-1)}\|_{H^s}\lesssim\delta,\\ &&u^{(m-1)}(t,x)|_{x\in\partial \Omega} = 0, \end{array} \right.

    and

    \left\{ \begin{array}{lll} &&d^{(m-1)}(0,x) = 0,\\ &&\|d^{(m-1)}\|_{H^s}\lesssim\delta,\\ &&d^{(m-1)}(t,x)|_{x\in\partial \Omega} = 0. \end{array} \right.

    Then we will find the m -th (m\geq2) approximation solution (u^{(m)}(t, \lambda x), d^{(m)}(t, \lambda x)) , which is equivalent to find (\textbf{h}^{(m)}, \textbf{w}^{(m)}) such that

    \begin{equation} \begin{aligned} &u^{(m)}(t,\lambda x) = u^{(m-1)}(t,\lambda x)+\textbf{h}^{(m)}(t,\lambda x),\\ &d^{(m)}(t,\lambda x) = d^{(m-1)}(t,\lambda x)+\textbf{w}^{(m)}(t,\lambda x). \end{aligned} \end{equation} (3.12)

    Substituting (3.12) into (2.3), we have

    \begin{aligned} \mathcal L_1(u^{(m)},d^{(m)})& = \mathcal L_1(u^{(m-1)},d^{(m-1)})+ \mathcal J_1[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)})+ \mathcal R_1(\textbf{h}^{(m)},\textbf{w}^{(m)}),\\ \mathcal L_2(u^{(m)},d^{(m)})& = \mathcal L_2(u^{(m-1)},d^{(m-1)})+ \mathcal J_2[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)})+ \mathcal R_2(\textbf{h}^{(m)},\textbf{w}^{(m)}). \end{aligned}

    Set

    \begin{aligned} & \mathcal J_1[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)}) = - \mathcal L_1(u^{(m-1)},d^{(m-1)}) = -E^{(m-1)},\\ & \mathcal J_2[u^{(m-1)},d^{(m-1)}](\textbf{h}^{(m)},\textbf{d}^{(m)}) = - \mathcal L_2(u^{(m-1)},d^{(m-1)}) = -\overline{E}^{(m-1)}, \end{aligned}

    which we supplement with the boundary conditions (3.3).

    Since we assume (u^{(m-1)}(t, \lambda x), d^{(m-1)}(t, \lambda x))\in \mathcal B_{{\varepsilon}} , there is the same structure between the linear system (2.12) and the linear system of m th approximation solutions. Thus, by the same arguments as in the proof of Proposition 2.1, we can show that above problem admits a solution (\textbf{h}^{(m)}, \textbf{w}^{(m)})\in H^s(\Omega)\times H^s(\Omega) . Here, we should use (2.2). Furthermore, similar to (2.61), we can use (3.9) and (3.10) to derive

    \|\textbf{h}^{(m)}\|^2_{H^s}+\|\textbf{w}^{(m)}\|^2_{H^s} \lesssim \|E^{(m-1)}\|^2_{H^s}+\|\overline{E}^{(m-1)}\|^2_{H^s},\quad \forall t > 0,

    where one can see the (m-1) -th error term (E^{(m-1)}, \overline{E}^{(m-1)}) such that

    \begin{aligned} E^{(m-1)}&: = \mathcal L_1(u^{(m-1)},d^{(m-1)}) = \mathcal R_1(\textbf{h}^{(m)},\textbf{w}^{(m)}),\\ \overline{E}^{(m-1)}&: = \mathcal L_2(u^{(m-1)},d^{(m-1)}) = \mathcal R_2(\textbf{h}^{(m)},\textbf{w}^{(m)}). \end{aligned}

    Moreover, by (3.5) and (3.6) and the standard Calderon-Zygmund theory, it follows that

    \|E^{(m)}\|_{H^s}+\|\overline{E}^{(m)}\|_{H^s} \lesssim N_m^3\Big(\|\textbf{h}^{(m)}\|^2_{H^s}+\|\textbf{w}^{(m)}\|^2_{H^s}+\|\textbf{w}^{(m)}\|^4_{H^s}\Big).

    The proof is now complete.

    Our target is to prove that the nonlinear Eq (1.1) admits global solutions

    (u^{(\infty)}(t,\lambda x),d^{(\infty)}(t,\lambda x)).

    This is equivalent to show that the series

    \begin{aligned} &\sum\limits_{i = 1}^m\textbf{h}^{(i)}(t,\lambda x) < +\infty,\\ &\sum\limits_{i = 1}^m\textbf{w}^{(i)}(t,\lambda x) < +\infty. \end{aligned}

    For a fixed constant s > 1 , let 1 < s = \bar{k} < k_0\leq k and

    \begin{aligned} &k_m: = \bar{k}+\frac{k-\bar{k}}{2^m},\quad k_{+\infty} = \bar{k},\\ &\alpha_{m+1}: = k_m-k_{m+1} = \frac{k-\bar{k}}{2^{m+1}}, \end{aligned}

    which yields

    \begin{equation} k_0 > k_1 > \ldots > k_m > k_{m+1} > \ldots. \end{equation} (4.1)

    Proposition 4.1. If the domain is the unbounded domain \mathbb R^+\times \mathbb R^2 , let the parameter \lambda > 1 . If the domain is a bounded domain \Omega , let the parameter \lambda = 1 and {\gamma\over{\varepsilon}^2} < 1 . Assume the initial approximation function (u^{(0)}, d^{(0)}) satisfying (2.4)–(2.8). Then the three dimension nematic liquid crystal flow (1.1) with the small initial data (1.4) and the boundary condition (1.5) admit a global Sobolev solution

    \begin{aligned} u^{(\infty)}(t,x)& = u^{(0)}(t,x)+\sum\limits_{m = 1}^{\infty}\boldsymbol{h}^{(m)}(t, \lambda x)+u_0(x)\in H^s(\Omega),\\ d^{(\infty)}(t,x)& = d^{(0)}(t,x)+\sum\limits_{m = 1}^{\infty}\boldsymbol{w}^{(m)}(t, \lambda x)+d_0(x)\in H^s(\Omega). \end{aligned}

    Moreover, it holds the estimate

    \|u^{(\infty)}\|_{H^s}+\|d^{(\infty)}\|_{H^s}\lesssim\delta.

    Proof. The proof is based on the induction. For convenience, we first deal with the case of zero initial data, that is, u_0(x) = \Big(0, 0, 0\Big)^T and d_0(x) = \Big(0, 0, 0\Big)^T . After that, we discuss the case of non-zero initial data. Note that N_m = N_0^m with N_0 > 1 . For all m = 1, 2, \ldots , we claim that there exists a sufficient small positive constant \delta such that

    \begin{equation} \begin{aligned} &\|\textbf{h}^{(m)}\|_{H^{k_{m-1}}}+\|\textbf{d}^{(m)}\|_{H^{k_{m-1}}} < \delta^{2^{m-1}},\\ &\|E^{(m)}\|_{H^{k_{m-1}}}+\|\overline{E}^{(m)}\|_{H^{k_{m-1}}} < \delta^{2^{m}},\\ &(u^{(m)},d^{(m)})\in \mathcal B_{\delta}. \end{aligned} \end{equation} (4.2)

    For the case of m = 1 , we recall the assumptions (2.4)–(2.8) on the initial approximation function (u^{(0)}(t, x), d^{(0)}(t, x)) . By (2.61), letting 0 < \delta_0 < N_0^{-(8+k-\bar{k})}\delta^2\ll1 , we obtain

    \|\textbf{h}^{(1)}\|_{H^{k_0}}+\|\textbf{w}^{(1)}\|_{H^{k_0}}\lesssim \|E^{(0)}\|_{H^{k_0}}+\|\overline{E}^{(0)}\|_{H^{k_0}} < \delta_0 < \delta^2.

    Moreover, by (3.8) and the above estimate,

    \begin{aligned} \|E^{(1)}\|_{H^{k_0}}+\|\overline{E}^{(1)}\|_{H^{k_0}}&\lesssim\| \mathcal R_1(\textbf{h}^{1)},\textbf{w}^{1)})\|_{H^{k_0}}+\| \mathcal R_2(\textbf{h}^{1)},\textbf{w}^{1)})\|_{H^{k_0}}\\ &\lesssim N_1^2\Big(\|\textbf{h}^{(1)}\|^2_{H^{k_0}}+\|\textbf{w}^{(1)}\|^2_{H^{k_0}}+\|\textbf{w}^{(1)}\|^4_{H^{k_0}}\Big)\\ &\lesssim N_1^2\Big(\|\textbf{h}^{(1)}\|_{H^{k_0}}+\|\textbf{w}^{(1)}\|_{H^{k_0}}\Big)^2\\ &\lesssim\delta^2, \end{aligned}

    and

    \|u^{(1)}\|_{H^{k_0}}+\|d^{(1)}\|_{H^{k_0}}\lesssim\|u^{(0)}\|_{H^{k_0}}+\|d^{(0)}\|_{H^{k_0}}+\|\textbf{h}^{(1)}\|_{H^{k_0}}+\|\textbf{w}^{(1)}\|_{H^{k_0}}\lesssim \delta,

    which means that (u^{(1)}, d^{(1)})\in \mathcal B_{\delta} .

    Assume that the case of m-1 holds, that is,

    \begin{equation} \begin{aligned} &\|\textbf{h}^{(m-1)}\|_{H^{k_{m}}}+\|\textbf{w}^{(m-1)}\|_{H^{k_{m}}} < \delta^{2^{m-2}},\\ &\|E^{(m-1)}\|_{H^{k_{m}}}+\|\overline{E}^{(m-1)}\|_{H^{k_{m}}} < \delta^{2^{m-1}},\\ &(u^{(m-1)},d^{(m-1)})\in \mathcal B_{\delta}. \end{aligned} \end{equation} (4.3)

    Then we prove that the case of m holds. Using (2.2), (3.7) and the second inequality of (4.3), we derive

    \begin{equation} \begin{aligned} \|\textbf{h}^{(m)}\|_{H^{k_{m-1}}}+\|\textbf{w}^{(m)}\|_{H^{k_{m-1}}}&\lesssim \|\Pi_{N_m}E^{(m-1)}\|_{H^{k_{m-1}}}+ \|\Pi_{N_m}\overline{E}^{(m-1)}\|_{H^{k_{m-1}}}\\ &\lesssim N_{m}^{\alpha_m}\Big(\|E^{(m-1)}\|_{H^{k_{m}}}+\|\overline{E}^{(m-1)}\|_{H^{k_{m}}}\Big)\\ & < \delta^{2^{m-2}}, \end{aligned} \end{equation} (4.4)

    which combined with (2.2), (3.8) and (4.1) yields

    \begin{equation} \begin{aligned} \|E^{(m)}\|_{H^{k_{m}}}+\|\overline{E}^{(m)}\|_{H^{k_{m}}} &\lesssim N_m^2\Big(\|\textbf{h}^{(m)}\|^2_{H^{k_{m}}}+\|\textbf{w}^{(m)}\|^2_{H^{k_{m}}}+\|\textbf{w}^{(m)}\|^4_{H^{k_{m}}}\Big)\\ &\lesssim N_{m}^{2+\alpha_{m+1}}\Big(\|E^{(m-1)}\|_{H^{k_{m+1}}}+\|\overline{E}^{(m-1)}\|_{H^{k_{m+1}}}\Big)^2\\ &\lesssim \ldots,\\ &\lesssim \Big[ N_0^{8+k-\bar{k}}\Big(\|E^{(0)}\|_{H^{k_{2m}}}+\|\overline{E}^{(0)}\|_{H^{k_{2m}}}\Big)\Big]^{2^m}. \end{aligned} \end{equation} (4.5)

    We choose a sufficiently small positive constant \delta_0 such that

    0 < N_0^{8+k-\bar{k}}\Big(\|E^{(0)}\|_{H^{k_{2m}}}+\|\overline{E}^{(0)}\|_{H^{k_{2m}}}\Big) < 2N_0^4\delta_0 < \delta^2.

    Thus, by (4.5) we have

    \|E^{(m)}\|_{H^{k_{m}}}+\|\overline{E}^{(m)}\|_{H^{k_{m}}} < \delta^{2^{m}},

    and

    \begin{aligned} 0&\leq \lim\limits_{m\rightarrow +\infty}\Big(\|E^{(m)}\|_{H^{k_{m}}}+\|\overline{E}^{(m)}\|_{H^{k_{m}}}\Big)\\ &\lesssim \Big[N_0^{8+k-\bar{k}}\Big(\|E^{(0)}\|_{H^{k_{+\infty}}}+\|\overline{E}^{(0)}\|_{H^{k_{+\infty}}}\Big)\Big]^{2^{+\infty}} \longrightarrow0. \end{aligned}

    So, the error term goes to 0 as m\rightarrow \infty , that is,

    \lim\limits_{m\rightarrow +\infty}(\|E^{(m)}\|_{H^{k_{m}}}+\|\overline{E}^{(m)}\|_{H^{k_{m}}}) = 0.

    On the other hand, note that N_m = N_0^m , by (4.3) and (4.4). It follows that

    \begin{aligned} \|u^{(m)}\|_{H^{k_{m}}}+\|d^{(m)}\|_{H^{k_{m}}}&\lesssim \|u^{(m-1)}\|_{H^{k_{m}}}+\|\textbf{h}^{(m)}\|_{H^{k_{m}}}+\|d^{(m-1)}\|_{H^{k_{m}}}+\|\textbf{w}^{(m)}\|_{H^{k_{m}}}\\ &\lesssim \delta+N_m^3\delta^{2^m}\lesssim\delta. \end{aligned}

    This means that (u^{(m)}, d^{(m)})\in \mathcal B_{\delta} . Hence we conclude that (4.2) holds.

    Therefore, the three dimension nematic liquid crystal flow (1.1) with the zero initial data admit global solutions

    \begin{aligned} u^{(\infty)}(t,x)& = u^{(0)}(t,x)+\sum\limits_{m = 1}^{\infty}\textbf{h}^{(m)}(t, \lambda x)+u_0(x) = u^{(0)}(t,x)+u_0(x)+ \mathcal O(\delta^2),\\ d^{(\infty)}(t,x)& = d^{(0)}(t,x)+\sum\limits_{m = 1}^{\infty}\textbf{w}^{(m)}(t, \lambda x)+d_0(x) = d^{(0)}(t,x)+d_0(x)+ \mathcal O(\delta^2). \end{aligned}

    Next, we discuss the case of small non-zero initial data.

    We introduce the auxiliary function

    \begin{aligned} \overline{u}(t,x)& = u(t,x)-u_0(x),\quad x\in\Omega,\\ \overline{d}(t,x)& = d(t,x)-d_0(x),\quad x\in\Omega, \end{aligned}

    Thus, the initial data reduces to

    \overline{u}(0,x) = \Big(0,0,0\Big),\quad \overline{d}(0,x) = \Big(0,0,0\Big),

    and the boundary condition (1.5) is changed into the Dirichlet boundary condition

    \overline{u}(t,x)|_{x\in\partial \Omega} = 0,\quad \overline{d}(t,x)|_{x\in\partial \Omega} = 0,

    and Eq (1.1) are transformed into equations of (\overline{u}, \overline{d}) .

    Thus, we can follow the above iteration scheme to construct a global Sobolev solution (\overline{u}, \overline{d}) . Furthermore, the global Sobolev solution of Eq (1.1) with a small non-zero initial data takes the form (\overline{u}(t, x)+u_0, \overline{d}(t, x)+d_0) . We notice that by the auxiliary function, the new first equation contain the non-autonomous force:

    f(u_0,d_0): = -\nu\Delta U_0+(U_0\cdot\nabla)U_0+\nabla P_0-\kappa\nabla\cdot(\nabla d_0\odot\nabla d_0),

    and the new second equation contain the non-autonomous force:

    g(u_0,d_0): = -\gamma\Delta d_0+u_0\cdot\nabla d_0-{\gamma\over{\varepsilon}^2}(|d_0|^2-1)d_0.

    Thus, we have to require the initial data (u_0, d_0)\in H^{s+2}(\Omega)\times H^{s+2}(\Omega) to be small and \nabla\cdot u_0 = 0 .

    Finally, using (1.2), by the standard Calderon-Zygmund theory for the Riesz operator \mathcal R , we have \| \mathcal R u\|_{ \mathbb L^{s_0}}\leq\|u\|_{ \mathbb L^{s_0}} with 1 < s_0 < \infty , hence

    \|P\|_{H^s}\lesssim\delta.

    This completes the proof.

    The authors declare that they have no competing interests.



    [1] F. Lin, C. Liu, Nonparabolic dissipative systems modeling the flow of liquid crystals, Commun. Pur. Appl. Math., 48 (1995), 501–537. http://dx.doi.org/10.1002/cpa.3160480503 doi: 10.1002/cpa.3160480503
    [2] F. Lin, C. Liu, Partial regularity of thed ynamic system modeling the flow of liquid crystals, Discrete Cont. Dyn.-A, 2 (1996), 1–22. http://dx.doi.org/10.3934/dcds.1996.2.1 doi: 10.3934/dcds.1996.2.1
    [3] J. Ericksen, Equilibrium theory of liquid crystals, In: Advances in liquid crystals, New York: Academic Press, 1976,233–298. http://dx.doi.org/10.1016/B978-0-12-025002-8.50012-9
    [4] F. Leslie, Theory of flow phenomemum in liquid crystal, In: Advances in liquid crystals, New York: Academic Press, 1979, 1–81. http://dx.doi.org/10.1016/B978-0-12-025004-2.50008-9
    [5] F. Lin, Nonlinear theory of defects in nematic liquid crystals; phase transition and flow phenomena, Commun. Pur. Appl. Math., 42 (1989), 789–814. http://dx.doi.org/10.1002/cpa.3160420605 doi: 10.1002/cpa.3160420605
    [6] C. Fefferman, Existence and smoothness of the Navier-Stokes equations, The millennium prize problems, 57 (2000), 67.
    [7] J. Leray, Sur le mouvement d'un liquide visqueux emplissant l'espace, Acta Math., 63 (1934), 193–248. http://dx.doi.org/10.1007/BF02547354 doi: 10.1007/BF02547354
    [8] L. Caffarelli, R. Kohn, L. Nirenberg, Partial regularity of suitable weak solutions of the Navier-Stokes equations, Commun. Pur. Appl. Math., 35 (1982), 771–831. http://dx.doi.org/10.1002/cpa.3160350604 doi: 10.1002/cpa.3160350604
    [9] F. Lin, A new proof of the Caffarelli-Kohn-Nirenberg theorem, Commun. Pur. Appl. Math., 51 (1998), 241–257. http://dx.doi.org/10.1002/(SICI)1097-0312(199803)51:3<241::AID-CPA2>3.0.CO;2-A doi: 10.1002/(SICI)1097-0312(199803)51:3<241::AID-CPA2>3.0.CO;2-A
    [10] T. Buckmaster, V. Vicol, Nonuniqueness of weak solutions to the Navier-Stokes equation, Ann. of Math. 189 (2019), 101–144. http://dx.doi.org/10.4007/annals.2019.189.1.3
    [11] O. Benslimane, A. Aberqi, J. Bennouna, Existence and uniqueness of weak solution of p(x)-Laplace in Sobolev spaces with variable exponents in complete manifolds, Filomat, 35 (2021), 1453–1463. http://dx.doi.org/10.2298/FIL2105453B doi: 10.2298/FIL2105453B
    [12] P. Constantin, C. Foiaş, Navier-Stokes equations, Chicago: University of Chicago Press, 1988.
    [13] C. Foiaş, O. Manley, R. Rosa, R. Temam, Navier-Stokes equations and turbulence, Cambridge: Cambridge University Press, 2001. http://dx.doi.org/10.1017/CBO9780511546754
    [14] S. Gala, On the improved regularity criterion of the solutions to the Navier-Stokes equations, Commun. Korean Math., 35 (2020), 339–345. http://dx.doi.org/10.4134/CKMS.c190019 doi: 10.4134/CKMS.c190019
    [15] M. Ragusa, On weak solutions of ultraparabolic equations, Nonlinear Anal.-Theor., 47 (2001), 503–511. http://dx.doi.org/10.1016/S0362-546X(01)00195-X doi: 10.1016/S0362-546X(01)00195-X
    [16] V. Scheffer, Boundary regularity for the Navier-Stokes equations in a half-space, Commun. Math. Phys., 85 (1982), 275–299. http://dx.doi.org/10.1007/BF01254460 doi: 10.1007/BF01254460
    [17] R. Temam, Navier-Stokes equations: theory and numerical analysis, North-Holland: Elsevier, 1979. http://dx.doi.org/10.1007/BF01254460
    [18] R. Temam, Navier-Stokes equations and nonlinear functional analysis, Philadelphia: Society for Industrial and Applied Mathematics, 1995. http://dx.doi.org/10.1137/1.9781611970050
    [19] F. Lin, J. Lin, C. Wang, Liquid crystal flows in two dimensions, Arch. Rational Mech. Anal., 197 (2010), 297–336. http://dx.doi.org/10.1007/s00205-009-0278-x doi: 10.1007/s00205-009-0278-x
    [20] M. Hong, Global existence of solutions of the simplified Ericksen-Leslie system in dimension two, Calc. Var., 40 (2011), 15–36. http://dx.doi.org/10.1007/s00526-010-0331-5 doi: 10.1007/s00526-010-0331-5
    [21] F. Lin, C. Wang, Global existence of weak solutions of the nematic liquid crystal flow in dimension three, Commun. Pur. Appl. Math., 69 (2016), 1532–1571. http://dx.doi.org/10.1002/cpa.21583 doi: 10.1002/cpa.21583
    [22] T. Huang, F. Lin, C. Liu, C. Wang, Finite time singularity of the nematic liquid crystal flow in dimension three, Arch. Rational Mech. Anal., 221 (2016), 1223–1254. http://dx.doi.org/10.1007/s00205-016-0983-1 doi: 10.1007/s00205-016-0983-1
    [23] F. Lin, C. Lin, Existence of solutions for the Ericksen-Leslie system, Arch. Rational Mech. Anal., 154 (2000), 135–156. http://dx.doi.org/10.1007/s002050000102 doi: 10.1007/s002050000102
    [24] X. Hu, D. Wang, Global solution to the three dimensional incompressible flow of liquid crystals, Commun. Math. Phys., 296 (2010), 861–880. http://dx.doi.org/10.1007/s00220-010-1017-8 doi: 10.1007/s00220-010-1017-8
    [25] C. Cavaterra, E. Rocca, H. Wu, Optimal boundary control of a simplified Ericksen-Leslie system for nematic liquid crystal flows in 2D, Arch. Rational Mech. Anal., 224 (2017), 1037–1086. http://dx.doi.org/10.1007/s00205-017-1095-2 doi: 10.1007/s00205-017-1095-2
    [26] W. Yan, The motion of closed hypersurfaces in the central force field, J. Differ. Equations, 261 (2016), 1973–2005. http://dx.doi.org/10.1016/j.jde.2016.04.020 doi: 10.1016/j.jde.2016.04.020
    [27] W. Yan, Dynamical behavior near explicit self-similar blow up solutions for the Born-Infeld equation, Nonlinearity, 32 (2019), 4682.
    [28] W. Yan, Nonlinear stability of explicit self-similar solutions for the timelike extremal hypersurfaces in \mathbb R^{1+3}, Calc. Var., 59 (2020), 124. http://dx.doi.org/10.1007/s00526-020-01798-2 doi: 10.1007/s00526-020-01798-2
    [29] W. Yan, B. Zhang, Long time existence of solution for the bosonic membrane in the light cone gauge, J. Geom. Anal., 31 (2021), 395–422. http://dx.doi.org/10.1007/s12220-019-00269-1 doi: 10.1007/s12220-019-00269-1
    [30] W. Yan, V. Rǎdulescu, Global small finite energy solutions for the incompressible magnetohydrodynamics equations in \mathbb R^+\times \mathbb R^2, J. Differ. Equations, 277 (2021), 114–152. http://dx.doi.org/10.1016/j.jde.2020.12.031 doi: 10.1016/j.jde.2020.12.031
    [31] X. Zhao, W. Yan, Existence of standing waves for quasi-linear Schrödinger equations on \mathbb T^n, Adv. Nonlinear Anal., 9 (2020), 978–993. http://dx.doi.org/10.1515/anona-2020-0038 doi: 10.1515/anona-2020-0038
    [32] J. Nash, The embedding for Riemannian manifolds, In: Annals of mathematics, Princeton: Princeton University Press, 1956, 20–63.
    [33] J. Moser, A rapidly converging iteration method and nonlinear partial differential equations I, Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 20 (1966), 265–315.
    [34] L. Hörmander, Implicit function theorems, Lectures at Stanford University, 1977.
    [35] S. Alinhac, Existence d'ondes de raréfaction pour des systèmes quasi-linéaires hyperboliques multidimensionnels, Commun. Part. Diff. Eq., 14 (1989), 173–230. http://dx.doi.org/10.1080/03605308908820595 doi: 10.1080/03605308908820595
    [36] V. Barbu, Nonlinear semigroups and differential equations in Banach spaces, Dordrecht: Springer, 1976.
    [37] V. Yudovich, The linearization method in hydrodynamical stability theory, Providence: American Mathematical Society, 1989.
    [38] A. Pazy, Semigroups of linear operators and applications to partial differential equations, New York: Springer-Verlag, 1983. http://dx.doi.org/10.1007/978-1-4612-5561-1
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1623) PDF downloads(96) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog