Loading [MathJax]/jax/output/SVG/jax.js
Research article

Optimal decay rates of higher–order derivatives of solutions for the compressible nematic liquid crystal flows in R3

  • Received: 30 November 2021 Revised: 03 January 2022 Accepted: 12 January 2022 Published: 18 January 2022
  • MSC : 35Q35, 35B40, 76A15

  • In this paper, we are concerned with optimal decay rates of higher–order derivatives of the smooth solutions to the 3D compressible nematic liquid crystal flows. The main novelty of this paper is three–fold: First, under the assumptions that the initial perturbation is small in HN–norm (N3) and bounded in L1–norm, we show that the highest–order spatial derivatives of density and velocity converge to zero at the L2–rates is (1+t)34N2, which are the same as ones of the heat equation, and particularly faster than the L2–rate (1+t)14N2 in [J.C. Gao, et al., J. Differential Equations, 261: 2334-2383, 2016]. Second, if the initial data satisfies some additional low frequency assumption, we also establish the lower optimal decay rates of solution as well as its all–order spatial derivatives. Therefore, our decay rates are optimal in this sense. Third, we prove that the lower bound of the time derivatives of density, velocity and macroscopic average converge to zero at the L2–rate is (1+t)54. Our method is based on low-frequency and high-frequency decomposition and energy methods.

    Citation: Zhengyan Luo, Lintao Ma, Yinghui Zhang. Optimal decay rates of higher–order derivatives of solutions for the compressible nematic liquid crystal flows in R3[J]. AIMS Mathematics, 2022, 7(4): 6234-6258. doi: 10.3934/math.2022347

    Related Papers:

    [1] Xiufang Zhao . Decay estimates for three-dimensional nematic liquid crystal system. AIMS Mathematics, 2022, 7(9): 16249-16260. doi: 10.3934/math.2022887
    [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, 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
    [4] 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
    [5] 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
    [6] 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
    [7] Junling Sun, Xuefeng Han . Existence of Sobolev regular solutions for the incompressible flow of liquid crystals in three dimensions. AIMS Mathematics, 2022, 7(9): 15759-15794. doi: 10.3934/math.2022863
    [8] 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
    [9] Yangyang Chen, Yixuan Song . Space-time decay rate of the 3D diffusive and inviscid Oldroyd-B system. AIMS Mathematics, 2024, 9(8): 20271-20303. doi: 10.3934/math.2024987
    [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
  • In this paper, we are concerned with optimal decay rates of higher–order derivatives of the smooth solutions to the 3D compressible nematic liquid crystal flows. The main novelty of this paper is three–fold: First, under the assumptions that the initial perturbation is small in HN–norm (N3) and bounded in L1–norm, we show that the highest–order spatial derivatives of density and velocity converge to zero at the L2–rates is (1+t)34N2, which are the same as ones of the heat equation, and particularly faster than the L2–rate (1+t)14N2 in [J.C. Gao, et al., J. Differential Equations, 261: 2334-2383, 2016]. Second, if the initial data satisfies some additional low frequency assumption, we also establish the lower optimal decay rates of solution as well as its all–order spatial derivatives. Therefore, our decay rates are optimal in this sense. Third, we prove that the lower bound of the time derivatives of density, velocity and macroscopic average converge to zero at the L2–rate is (1+t)54. Our method is based on low-frequency and high-frequency decomposition and energy methods.



    In this paper, we investigate the upper and lower bounds of decay rates for global solution to compressible nematic liquid crystal flows in three–dimensional whole space:

    {ρt+div(ρu)=0,(ρu)t+div(ρuu)μΔu(μ+ν)divu+P(ρ)=γdΔd,dt+ud=θ(Δd+|d|2d), (1.1)

    where t0 is time and xR3 is the spatial coordinate. Here, the symbol is the Kronecker tensor product. We denote the unknown functions ρ(x,t) denotes the fluid density, u=(u1,u2,u3) is the fluid velocity, d is the macroscopic average of the nematic liquid crystal orientation field. The pressure P=P(ρ)=aργ(a>0,γ1) is a smooth function in a neighborhood of 1 with P(1)=1. μ and ν are shear viscosity and the bulk viscosity coefficients of the fluid, respectively, which satisfy the physical assumptions:

    μ>0,2μ+3ν0.

    The positive constants γ and θ represent the competition between the kinetic energy and the potential energy, and the microscopic elastic relaxation time for the molecular orientation field, respectively. For the sake of simplicity, we set the constants γ and θ to be 1. We consider the Cauchy problem of the system (1.1) subject to the initial conditions:

    (ρ,u,d)(x,t)t=0=(ρ0,u0,d0)(x)(1,0,ω0),as|x|, (1.2)

    where ω0 is a unit constant vector.

    Liquid crystal has important physical and chemical properties such as photoelectric effect, thermal effect, photochemical effect and so on. After nearly a century of research, liquid crystals have been widely used in production, life, and scientific research. Particularly, liquid crystal displays have been widely used in LED technology, airplanes, medical equipment, bioengineering, machinery manufacturing, and other fields. Next, let us present some explanations about the above model. The nematic liquid crystal flows are a coupling between the compressible Navier–Stokes equations and the transported flow harmonic maps. Ericksen [4] and Leslie [20] first established the continuum theory of liquid crystals in the 1960s. Since then, due to the physical importance and mathematical challenges, the study of the full Ericksen–Leslie model has attracted many physicists and mathematicians. Considering the compressible liquid crystal flows, Ding–Lin–Wang [2] gained both existence and uniqueness of global strong solution in one–dimensional space. For the case of multi-dimensional space, Jiang–Jiang–Wang in [16,17] proved the global existence of weak solutions to the initial–boundary problem with large initial energy. Huang–Wang–Wen in [13] obtained the blow up criterion of strong solutions. Hu–Wu [11] showed the existence and uniqueness of global strong solution in critical Besov spaces. Under the assumption that the initial energy is suitably small, Wu–Tan in [32] proved the global existence of small energy weak solution. Li–Xu–Zhang in [22] established the global existence of classical solution with smooth initial data which are of small energy but possibly large oscillations. Recently, Gao–Tao–Yao [8] obtained the global well–posedness of classical solution under the condition that the initial data is a small perturbation of the constant equilibrium state in the HN(R3)(N3)–framework. Furthermore, if the initial perturbation data belongs to L1 additionally, they obtained the optimal decay rate of k–th (kN1) order spatial derivative of solution in [8] as follow:

    k(ρ1)(t)HNk+ku(t)HNkC(1+t)34k2,k[0,N1], (1.3)
    l(dω0)(t)L2C(1+t)34l2,l[0,N+1]. (1.4)

    Here, the positive constant C is independent of time.

    When the director is a constant vector, then the compressible nematic liquid crystal flow (1.1) becomes the compressible Navier–Stokes equation. There are a lot of basic results about the global existence, unique and time decay rates of the solutions to the compressible Navier–Stokes equations, cf.[3,6,7,10,14,23,24,28,29] and references cited therein. We also mention that the problem of time-asymptotic behavior for the solutions of hydrodynamic equations is a hot topic, see [5,12,15,18,30,31] and references cited therein.

    By observing the results of [8], decay rate (1.3) shows that the highest–order spatial derivative of density and velocity converge to zero at L2–rate (1+t)14N2, which are slower than the L2–rate (1+t)34N2 for the heat equation, and thus are not optimal in this sense. Thus, we are caught up in the upper and lower optimal decay rates of higher–order spatial derivatives of the solutions to the 3D compressible liquid crystal flows (1.1). More precisely, we focus on the following three problems:

    (ⅰ) Can we show that the highest order spatial derivative of the density and velocity converge to zero at the same L2–rate (1+t)34N2 as that of the heat equation?

    (ⅱ) Can we provide some lower bounds of decay rate for the solution as well as its all–order spatial derivatives?

    (ⅲ) Can we provide some information on the upper and lower bound of decay rate for the time derivatives of solution?

    The main purpose of this article is to give a clear answer to the above three problems.

    In this article, we use Hk(R3) to denote the usual Sobolev spaces with norm Hk. Generally, we use Lp, 1p to denote the usual Lp(R3) spaces with norm Lp. The notation ab means that aCb for a universal positive constant which is independent of time t. Let Λs be the pseudo–differential operator defined by

    Λsf=F1(|ξ|sˆf)  forsR,

    where ˆf and F are the Fourier transform of f. For a radial function ϕC0(R3ξ) such that ϕ(ξ)=1 when |ξ|1 and ϕ(ξ)=0 when |ξ|2, we define the low–frequency part and the high–frequency part of f as follows

    fl=F1[ϕ(ξ)ˆf],andfh=F1[(1ϕ(ξ))ˆf].

    Before stating our main results, let us recall the result of [8] in the following.

    Theorem 1.1.(See [8]) Assume that the initial data (ρ01,u0,d0)HN(R3)L1(R3) for any integer N3, d0(x)=1 in R3 and there exists a small constant δ0>0 such that

    (ρ01,u0,d0)H3δ0, (1.5)

    then the Cauchy problem (1.1)–(1.2) admits a unique globally classical solution (ρ,u,d) such that for any t[0,),

    (ρ1,u,d)(t)2HN+t0(ρ2HN1+(u,2d)2HN)dτC(ρ01,u0,d0)2HN. (1.6)

    If the initial data d0ω0L2 and (ρ01,u0,d0)L1 are finite additionally, the global solution (ρ,u,d) of problem (1.1)–(1.2) satisfies for all t0

    k(ρ1)(t)HNk+ku(t)HNkC(1+t)34k2,k[0,N1],l(dω0)(t)L2C(1+t)34l2,l[0,N+1]. (1.7)

    Here, the positive constant C is independent of time.

    Now, we are in a position to state our main results, which are stated in the following four theorems. First, we show that the highest–order spatial derivative of the density and velocity converge to zero at the same L2–rate as ones of the heat equation.

    Theorem 1.2. Under the assumptions of Theorem 1.1, for all t0, it holds that

    N(ρ1)(t)L2+Nu(t)L2C(1+t)34N2. (1.8)

    Here, the positive constant C is independent of time.

    Next, we address the lower bound of decay rate for the solution as well as its all–order spatial derivatives.

    Theorem 1.3. Assume that all the hypotheses of Theorem 1.1 are in force, denote m0=ρ0u0, b0=Λd0, and the Fourier transform (^ρ01,ˆm0,^b0) satisfies

    |^ρ01(ξ)|c0,|ˆm0(ξ)|=0,|^b0(ξ)|c0,for0|ξ|1, (1.9)

    where c0 is a positive constant. Then, the global solution (ρ,u,d) has the decay rates for all tt

    min{k(ρ1)(t)L2,ku(t)L2}c1(1+t)34k2,k[0,N],l+1(dω0)(t)L2c2(1+t)34l2,l[0,N]. (1.10)

    Here, t is a positive large time, the two positive constants c1 and c2 are independent of time.

    Next, we will establish the upper bounds of decay rates for the time derivatives of the solution to the 3D compressible liquid crystal flows (1.1).

    Theorem 1.4. Under the assumptions of Theorem 1.1 and Theorem 1.2, the global solution (ρ,u,d) of Cauchy problem (1.1)–(1.2) satisfies for all t0

    ktu(t)L2C(1+t)54k2,k[0,N2],ltρ(t)L2C(1+t)54l2,lt(dω0)(t)L2C(1+t)74l2,l[0,N1]. (1.11)

    Here, the positive constant C is independent of time.

    Finally, we will establish the lower bounds of decay rates for the time derivatives of the solution to the 3D compressible liquid crystal flows (1.1).

    Theorem 1.5. Under all the assumptions in Theorem 1.1, and the condition (1.9) holds, then the global solution (ρ,u,d) of Cauchy problem (1.1)–(1.2) satisfies for all tt

    tu(t)L2c3(1+t)54,t(dω0)(t)L2c3(1+t)54. (1.12)

    Furthermore, if there exists a small constant δ1 such that u0L1δ1, it holds that for all tt

    min{tϱ(t)L2,divu(t)L2}c3(1+t)54. (1.13)

    Here, t is a positive large time, and the positive constant c3 is independent of time.

    Remark 1.6. Compared to Theorem 1.1 of [8], the main innovation of Theorem 1.2–1.5 lies in the following three aspects: First, by observing the decay rates in (1.8), we prove that the highest–order spatial derivative of the density and velocity converge to their corresponding equilibrium states at the L2–rate (1+t)34N2, which is the same as one of the heat equation and particularly faster than the rates (1+t)14N2 in [8]. Second, for well–chosen initial data, Theorem 1.3 also gives the lower bounds on solution as well as its all–order spatial derivatives. Thus, our time decay rates are really optimal in this sense. Third, we also gives the lower bound of decay rates for the time derivatives of density, velocity and macroscopic average for the 3D compressible liquid crystal flows, which converge to zero at the L2–rates (1+t)54, (1+t)54 and (1+t)54, respectively.

    Now, let us sketch the main strategy of proving Theorem 1.2–Theorem 1.5 and explain some main difficulties and techniques involved in the process. Roughly speaking, we will make full use of the benefit of the low–frequency and high–frequency decomposition f=fl+fh, where fl and fh stand for the low–frequency part and high–frequency part of f, respectively.

    For the proof of Theorem 1.2, motivated by the work in [27], we will establish the highest–order spatial derivatives of the density and velocity to compressible nematic liquid crystal flows (1.1). There are four steps to achieve this. First, we rewrite the Cauchy problem (1.1)–(1.2) into the system (2.1). We notice that the low-frequency part k(ϱ,u)(0kN) of the corresponding linear system to (2.1) has been obtained by [21]. Then, by employing Duhamel's principle, the key linear decay estimates in Lemma 3.1 and nonlinear energy estimates, we can get the optimal decay rate of N(ϱl,ul)(t)L2 (see the proof of Lemma 3.3 for details). Second, we want to obtain the optimal decay rate of N(ϱh,uh)(t)L2. Through high–frequency and low–frequency decomposition and precise energy estimates, we establish the energy inequality as follows:

    12ddtR3|Nϱh|2+|Nuh|2dx+C1N+1uh2L2(1+t)32N  +(δ+(1+t)32)N+1u2L2+δN(ϱ,u)2L2. (1.14)

    Third, note that the energy equality (1.14) only gives the dissipation estimate for uh. In order to explore the dissipation estimates for ϱh, we will construct the new interactive energy functionals between uh and ϱh. Therefore, we have

    ddtR3N1uhNϱhdx+C2Nϱh2L2(1+t)32N+(δ+(1+t)32)Nϱ2L2  +(1+(1+t)32)Nu2L2+(1+δ)N+1u2L2. (1.15)

    Fourth, we choose two sufficiently large positive constant D0 and T0, then define the temporal energy functional as

    E(t)=D0N(ϱh,uh)2L2+R3N1uhNϱhdx. (1.16)

    Notice that E(t) is equivalent to N(ϱh,uh)2L2. Multiplying (1.14) by D0, adding the resulting inequality with (1.15), and then for all tT0, we obtain the Lyapunov-type energy inequality as follow:

    ddtE(t)+C3E(t)(1+t)32N+Nϱl2L2+Nul2H1. (1.17)

    By virtue of Lemma 3.3 and Gronwall's inequality, we obtain the optimal L2 time decay rates of N(ϱh,uh)2L2. In addition, together with the N–order low–frequency decay rates of (3.7), we obtain the decay rates of (1.8) immediately. Thus, we have completed the proof of Theorem 1.2.

    For Theorem 1.3, we show the lower bounds on the decay rates of solutions. Compared to the proof of Theorem 1.2, the new difficulty we encounter is that the system (2.1) is not a conservative form, which implies that it seems impossible to obtain the lower optimal decay rates as in (1.10). To this end, let's break it down into two parts. For the part of Navier–stokes system, the key idea here is that, instead of using the system (2.1)1–(2.1)2, we will employ the system (4.1) of (ϱ,m) with m=(ϱ+1)u, which can be rewritten in the conservative form. As a result, one can shift the derivative onto the solution semigroup to obtain the desired lower optimal decay rates (See the proof of (4.4) for details). For the part of macroscopic average, we notice that the linearized system of (4.6) is mere heat equation on n. When the initial data satisfy (1.9), we will employ Plancherel theorem and careful analysis on the solution semigroup to obtain an optimal lower bound estimate for the linear part. We find a similar structure between (4.6) and the system (4.1), so we can prove the lower bound on the convergence rates in L2–norm for the macroscopic average(see the proof of Lemma 4.2 for details). Thus, we complete the proof of Theorem 1.3.

    For Theorem 1.4 and Theorem 1.5, we establish the upper and lower bound of decay rate for the time derivative of solution in L2–norm. It is worth mentioning that the lower decay rate estimate for the time derivative of solution can be obtained which is inspired by the work of Gao–Lyu–yao in [9]. However, the lower bound of decay estimates in [9] are established for the compressible fluid model of Korteweg type.

    In this section, we will reformulate the problem firstly. Set ϱ=ρ1 and n=dω0, the Cauchy problem (1.1)–(1.2) can be reformulated into:

    {ϱt+div(u)=S1,utμΔu(μ+ν)divu+ϱ=S2,ntΔn=S3,(ϱ,u,n)(x,t)t=0=(ϱ0,u0,n0)(x)(0,0,0),as|x|. (2.1)

    Here Si(i=1,2,3) are defined as

    S1:=ϱdivuuϱ,S2:=uuh(ϱ)[μΔu+(μ+ν)divu]f(ϱ)ϱg(ϱ)nΔn,S3:=un+|n|2(n+ω0), (2.2)

    where the three nonlinear functions of ϱ are defined by

    h(ϱ):=ϱϱ+1,f(ϱ):=P(ϱ+1)ϱ+11,andg(ϱ):=1ϱ+1. (2.3)

    Assume there exists a small positive constant δ satisfying following estimate

    (ϱ,u,n)(t)H3δ, (2.4)

    for all t[0,T]. By virtue of (2.4) and Sobolev inequality, it is easy to get

    12ϱ+132.

    Hence, we immediately have

    |h(ϱ)|,|f(ϱ)|C|ϱ|and|gk1(ϱ)|,|hk(ϱ)|,|fk(ϱ)|Ck1. (2.5)

    And, for the nonlinear terms of the model (2.1), employing the Hölder's inequality, we obtain

    (S1,S2)L1(ϱL2+uL2+nL2)(ϱL2+uH1+nH1)δ(ϱL2+uH1+nH1). (2.6)

    Let us to consider the following linearized compressible nematic liquid crystal system:

    {ϱt+div(u)=0,utμΔu(μ+ν)divu+ϱ=0,ntΔn=0, (3.1)

    with the initial conditions:

    (ϱ,u,n)(x,t)t=0=(ϱ0,u0,n0)(x)(0,0,0),as|x|.

    Since the system (3.1) is a decoupled system of the classical linearized Navier–Stokes equations and heat equations. If we set

    U(t)=(ˉϱ(t),ˉu(t))t,U(0)=(ˉϱ(0),ˉu(0))t.

    Then the solution to (3.1)1–(3.1)2 can be written as

    U(t)=etAU(0), (3.2)

    where A is a matrix–valued differential operators of the form

    A=(0divμΔ(μ+ν)div).

    The solution semigroup eAt has the following lemma on the decay in time.

    Lemma 3.1. Let k32 and 2r, then for any t0, it holds that

    ketAUl(0)LrC(1+t)34k2U(0)L1. (3.3)

    Moreover, if initial data satisfy (1.9), then there exists a positive large time tsuch that for all tt, we have

    min{kˉϱl(t)L2,kˉml(t)L2}c4(1+t)34k2, (3.4)

    for k=0,1.Here, the positive constant c4 is independent of time.

    Proof. The proof can be seen in [21].

    To treat the macroscopic average, we notice that the Eq (3.1)3 is a mere heat equation on n. We state the large-time behavior of solutions to the heat equation as the following lemma which can be obtained by direct calculation or more can refer to [26,33].

    Lemma 3.2. Let k32 and 2r, then for any t0, it holds that

    k(ˉn)l(t)LrC(1+t)34k2ˉn(0)L1. (3.5)

    Moreover, if initial data satisfy (1.9), then there exists a positive large time t, such that for all tt, we have

    ˉnl(t)L2c4(1+t)34. (3.6)

    Here, the positive constant c4 is independent of time.

    Second, we state the L2–time decay rates on the low–frequency part of the solution in the nonlinear system (2.1).

    Lemma 3.3. Under the assumptions of Theorem 1.1, then solution (ϱ,u) of the nonlinear system (2.1) satisfiesthe following decay rates:

    N(ϱl,ul)(t)L2(1+t)34N2. (3.7)

    Proof. We write the nonlinear terms S:=( S1, S2)T. By virtue of the Eq (2.1)1–(2.1)2, Lemma 3.1, Duhamel's principle, Plancherel theorem and Hausdorff–Young's inequality, we have

    N(ϱl,ul)(t)L2(1+t)34N2(ϱ,u)(0)L1+t20(1+tτ)34N2S(τ)L1dτ+tt2(1+tτ)54|ξ|N1ˆSl(τ)Ldτ. (3.8)

    On the other hand, by employing decay rate (1.7), Lemma 7.1, Lemma 7.3 and Hölder's inequality, we can bound the third term on the right–hand of (3.8) as follows

    |ξ|N1ˆSl(τ)LN2S(t)lL1N2(ϱu)L1+N2(ϱu)L1+N2(uu)L1+N2(h(ϱ)Δu)L1+N2(h(ϱ)divu)L1+N2(f(ϱ)ϱ)L1+N2(g(ϱ)nΔn)L1(ϱ,u)L2N2(ϱ,u)L2+(ϱ,u)L2N1(ϱ,u)L2+ϱL2NuL2+2uL2N2ϱL2+ϱL2N2(nΔn)L2+nΔnL2N2ϱL2(1+t)54(1+t)2N14+(1+t)34(1+t)2N+14+(1+t)34(1+t)2N+34+(1+t)74(1+t)2N14+(1+t)34(1+t)5+N2+(1+t)72(1+t)2N14(1+t)1N2. (3.9)

    Here, for the term nΔnL2, we have

    nΔnL2nL3ΔnL6nH13nL2(1+t)54(1+t)94(1+t)72. (3.10)

    For the term N2(nΔn)L2, we get

    N2(nΔn)L2nLNnL2+ΔnL3N1nL62nH1NnL2(1+t)74(1+t)3+2N4(1+t)52N2. (3.11)

    Substituting (3.9) and (2.6) into (3.8), we can get the follow estimates

    N(ϱl,ul)(t)L2C(1+t)34N2+tt2(1+tτ)54(1+t)1N2dτ.(1+t)34N2. (3.12)

    Hence, we complete the proof of this lemma.

    Third, we will give the energy estimates which contains the dissipation of Nuh.

    Lemma 3.4. Under the assumptions of Theorem 1.1, then we have

    12ddtR3|Nϱh|2+|Nuh|2dx+C1N+1uh2L2(1+t)32N  +(δ+(1+t)32)N+1u2L2+δN(ϱ,u)2L2, (3.13)

    for any t[0,).

    Proof. Taking

    F1(1ϕ(ξ))N(2.1)1,Nϱh+F1(1ϕ(ξ))N(2.1)2,Nuh, (3.14)

    and using integration by parts, we can obtain

    12ddtR3|Nϱh|2+|Nuh|2dx+(2μ+ν)N+1uh2L2=NSh1,Nϱh+NSh2,Nuh=I1+I2. (3.15)

    The right-hand side of (3.15) can be estimated one by one. For the term I1, it holds that

    I1=NSh1,Nϱh=N(uϱ)h,NϱhN(ϱu)h,Nϱh=I11+I12. (3.16)

    The first term I11 can be rewritten as follows

    I11=N(uϱ)h,Nϱh=N(uϱ)N(uϱ)l,Nϱh=N(uϱh)+N(uϱl)N(uϱ)l,Nϱh=I111+I112+I113. (3.17)

    For the term I111, due to Lemma 7.4, Hölder's inequality, Young's inequality and Sobolev interpolation theorem, we arrive at

    |I111||uN+1ϱh,Nϱh|+|[N,u]ϱh,Nϱh|12divu,(Nϱh)2+[N,u]ϱhL2NϱhL2uLNϱ2L2+(uLNϱhL2+NuL2ϱhL)Nϱ2L22uH1Nϱ2L2+(2uH1NϱhL2+2ϱH1NuL2)Nϱ2L2δN(ϱ,u)2L2. (3.18)

    Here we have defined the commutator:

    [N,u]ϱh=N(uϱh)uN+1ϱh.

    For the term I112, we can obtain

    |I112|=|N(uϱl),Nϱh|N(uϱl)L2NϱhL2(uLN+1ϱlL2+ϱlL3NuL6)NϱhL2(uH2NϱL2+ϱH1N+1uL2)NϱhL2δN+1u2L2+δNϱ2L2. (3.19)

    Similarly, it is easy to see that

    |I113|=|N(uϱ)l,Nϱh|N1(uϱ)L2NϱhL2(uLNϱL2+ϱL3N1uL6)NϱhL2(uH2NϱL2+ϱH1NuL2)NϱhL2δN(ϱ,u)2L2. (3.20)

    Substituting (3.18)–(3.20) into (3.17), we can conclude that

    |I11|δNu2H1+δNϱ2L2. (3.21)

    For the term I12, by using the Lemma 7.3, Hölder's inequality, Young's inequality, Sobolev interpolation theorem, we have

    |I12|N(ϱu)hL2NϱhL2N(ϱu)L2NϱhL2(ϱLNuL2+uLNϱL2)NϱhL2(ϱH2N+1uL2+uH2NϱL2)NϱhL2δNϱ2L2+δN+1u2L2. (3.22)

    For the term I2, it holds that

    I2=N(uu)h,Nuh+N(h(ϱ)Δu)h,Nuh+N(h(ϱ)(divu))h,Nuh+N(f(ϱ)ϱ)h,Nuh+N(g(ϱ)nΔn)h,Nuh=5i=1I2i. (3.23)

    For the term I21, making use of integration by parts, we have

    |I21|=|N1(uu)h,Nuh|N1(uu)L2NuhL2(uLNuL2+uL3N1uL6)NuhL2uH1NuL2NuL2(1+t)54(1+t)14N2N+1uL2(1+t)32N+(1+t)32N+1u2L2. (3.24)

    For the term I22, we obtain

    |I22|=|N1(h(ϱ)Δu)h,Nuh|N1(h(ϱ)Δu)L2NuhL2(ρLN+1uL2+2uL3N1ϱL6)NuhL2(ϱH1N+1uL2+2uH1NϱL2)NuhL2δN+1u2L2+(1+t)74(1+t)14N2N+1uL2(1+t)32N+(δ+(1+t)52)N+1u2L2. (3.25)

    Similarly, it is easy to see that

    |I23|=|N1(h(ϱ)(divu))h,Nuh|N1(h(ϱ)(divu))L2NuhL2(ρLN+1uL2+2uL3N1ϱL6)NuhL2(ϱH1N+1uL2+2uH1NϱL2)NuhL2δN+1u2L2+(1+t)74(1+t)14N2N+1uL2(1+t)32N+(δ+(1+t)52)N+1u2L2. (3.26)

    For the term I24, similar to the proof of (3.24), we have

    |I24|=|N1(f(ϱ)ϱ)h,Nuh|N1(f(ϱ)ϱ)L2NuhL2(ϱLNϱL2+ϱL3N1ϱL6)NuhL2ϱH1NϱL2NuL2(1+t)54(1+t)14N2N+1uL2(1+t)32N+(1+t)32N+1u2L2. (3.27)

    For the term I25, by virtue of Lemma 7.3, we have

    |I25|=|N1(g(ϱ)nΔn)h,Nuh|N1(g(ϱ)nΔn)L2NuhL2(g(ϱ)LN1(nΔn)L2+nΔnL3N1g(ϱ)L6)NuhL2(ϱH1N1(nΔn)L2+nΔnL3NϱL2)NuL2[ϱH1(nLN+1nL2+ΔnL3NnL6)+(nL6ΔnL6)NϱL2]NuL2ϱH12nH1N+1nL2NuL2+2nL23nL2NϱL2NuL2(1+t)54(1+t)74(1+t)54N2NuL2+δNϱL2NuL2(1+t)32N+(δ+(1+t)7)N+1u2L2+δNϱ2L2. (3.28)

    Substituting (3.24)–(3.28) into (3.23), we have

    |I2|(1+t)32N+(δ+(1+t)32)N+1u2L2+δNϱ2L2. (3.29)

    Substituting (3.21)–(3.22) and (3.29) into (3.15), we obtain the estimate (3.13).

    Next, we will establish the dissipation estimates for Nϱh.

    Lemma 3.5. Suppose that the assumptions of Theorem 1.1 are in force, then we have

    ddtR3N1uhNϱhdx+C2Nϱh2L2(1+t)32N+(δ+(1+t)32)Nϱ2L2  +(1+(1+t)32)Nu2L2+(1+δ)N+1u2L2, (3.30)

    for any t[0,).

    Proof. Taking F1(1ϕ(ξ))N1(2.1)2,Nϱh, and making use of the integration by parts, it hold that

    R3N1uhtNϱhdx+P(1)Nϱh2L2=μN1Δuh,Nϱh+(μ+ν)Ndivuh,Nϱh+N1Sh2,Nϱh. (3.31)

    In order to deal with the term R3N1uhtNϱhdx, we apply the transport equation (2.1)1. More precisely, we have

    R3N1uhtNϱhdx=ddtR3N1uhNϱhdxR3N1uhNϱhtdx=ddtR3N1uhNϱhdx+R3N1divuhN1ϱhtdx=ddtR3N1uhNϱhdx+R3N1divuhN1(divuS1)hdx. (3.32)

    Substituting (3.33) into (3.32), and employing Young's inequality, it is easy to deduce

    ddtR3N1uhNϱhdx+C2Nϱh2L2Nuh2L2+N+1uh2L2+N1Sh2,NϱhR3N1divuhN1Sh1dx. (3.33)

    Here, we write

    Π1+Π2=R3N1divuhN1Sh1dx+N1Sh2,Nϱh.

    For the term Π1, with the help of Hölder's inequality and Lemma 7.3, it holds that

    |Π1|=|N(ϱu)h,Nuh|N(ϱu)L2NuhL2(ϱ,u)LN(ϱ,u)L2NuhL2(ϱ,u)H1N(ϱ,u)L2NuhL2(1+t)54(1+t)14N2NuhL2(1+t)32N+(1+t)32Nuh2L2. (3.34)

    For the term Π2, similarly to the proof of (3.16), we have

    Π2=N1(uu)h,Nϱh+N1(h(ϱ)Δu)h)h,Nϱh+N1(h(ϱ)(divu))h,Nϱh+N1(f(ϱ)ϱ)h,Nϱh+N1(g(ϱ)nΔn)h,Nϱh:=5i=1Π2i. (3.35)

    For the term Π21, we have

    |Π21|N1(uu)L2NϱhL2(uLNuL2+uL3N1uL6)NϱhL2u|H1NuL2NϱhL2(1+t)54(1+t)14N2NϱhL2(1+t)32N+(1+t)32NϱhL2. (3.36)

    For the term Π22, it easy to deduce

    |Π22|N1(h(ϱ)Δu)L2k+1ϱhL2(ϱLN+1uL2+2uL3N1ϱL6)NϱhL2(ϱH2N+1uL2+2uH1NϱL2)NϱL2δN+1u2L2+δNϱ2L2. (3.37)

    For the term Π23, we can get

    |Π23|N1(h(ϱ)(divu))L2k+1ϱhL2(ϱLN+1uL2+2uL3N1ϱL6)NϱhL2(ϱH2N+1uL2+2uH1NϱL2)NϱL2δN+1u2L2+δNϱ2L2. (3.38)

    For the term Π24, similarly to the proof of (3.34), we have

    |Π24|N1(f(ϱ)ϱ)L2NϱhL2(ϱLNϱL2+ϱL3N1ϱL6)NϱhL2ϱH1NϱL2NϱhL2(1+t)54(1+t)14N2NϱhL2(1+t)32N+(1+t)32NϱhL2. (3.39)

    For the term Π25, by virtue of Lemma 7.3, it is easy to obtain

    |Π25|N1(g(ϱ)nΔn)L2NϱhL2(g(ϱ)LN1(nΔn)L2+nΔnL3N1g(ϱ)L6)NϱhL2ϱL(nLN+1nL2+ΔnL3NnL6)NϱhL2+nL6ΔnL6Nϱ2L2ϱH12nH1N+1nL2NϱL2+2nL23nL2Nϱ2L2(1+t)54(1+t)74(1+t)54N2NϱL2+δNϱ2L2(1+t)32N+(δ+(1+t)7)Nϱ2L2. (3.40)

    Substituting (3.36)–(3.41) into (3.35), yields directly

    |Π2|(1+t)32N+(δ+(1+t)32)Nϱ2L2+δN+1u2L2. (3.41)

    Substituting (3.34) and (3.41) into (3.33), we complete the proof of the Lemma 3.5.

    Finally, we will close the estimate by combining them which have been already proved along the proof of Lemma (3.3)–(3.5). To achieve this, we choose sufficiently large time T0 and positive constant D0, and then define the temporary energy functional

    E(t)=D0N(ϱh,uh)2L2+R3N1uhNϱhdx,

    for tT0, it is noticed that E(t) is equivalent to N(ϱh,uh)2L2. Substituting (3.13) and (3.30) into

    D0×(3.13)+(3.30), (3.42)

    which together with the smallness of δ, for tT0, it holds that

    ddtE(t)+Nϱh2L2+N+1uh2L2(1+t)3+2N2+Nϱl2L2+Nul2H1, (3.43)

    where we have used the fact that T0 is large enough. On other hand, it is clear that

    Nϱh2L2+N+1uh2L2C3E(t).

    Hence, by virtue of estimate (3.7) and Gronwall's inequality, we can arrive at

    N(ϱh,uh)2L2(1+t)32N. (3.44)

    So, the proof of Theorem 1.2 has been completed.

    In this section, we will establish the lower optimal decay rates of the solution as well as its all–order spatial derivatives in Theorem 1.3. At first, we rewrite (1.1)1–(1.1)2 with m=(ϱ+1)u in the perturbation form as follow:

    {ϱt+divm=0,mt+P(1)ϱμΔm(μ+ν)divm=divF, (4.1)

    where the functions F=F(ϱ,u,n) is defined as

    F=[P(ϱ+1)P(1)ϱ]I3×3+μ(ϱu)+(μ+ν)div(ϱu)I3×3+(1+ϱ)(uu)+12(|n|2)I3×3(nn).

    Here nn=(in,jn)1i,j3. It is easy to check

    nΔn=div(nn)12(|n|2), (4.2)

    with the initial conditions:

    (ϱ,m)(x,t)t=0=(ϱ0,m0)(x)(0,0),as|x|.

    Next, we will establish the lower bound decay rate of k(ϱ,u) to the system (2.1).

    Lemma 4.1. Under all the assumptions of Theorem 1.3, then the global solution (ϱ,u) has the decay rates for all tt and 0kN

    min{kϱ(t)L2,ku(t)L2}c1(1+t)34k2. (4.3)

    Here, t is a positive large time, the positive constant c1 is independent of time.

    Proof. By virtue of Duhamel's principle, system (4.1), Lemma 3.1, Theorem 1.1 for k=0,1, we have

    min{kϱ(t)L2,ku(t)L2}min{kϱ(t)L2,km(t)L2}min{kϱl(t)L2,kml(t)L2}c4(1+t)34k2t0(1+tτ)54k2FL1dτc4(1+t)34k2t0(1+tτ)54k2(1+τ)32dτc4(1+t)34k2C(1+t)54k2c1(1+t)34k2. (4.4)

    Then, for 2kN, by applying (4.3) and Lemma 7.2, we have from Sobolev interpolation that

    k(ϱ,u)(t)L2C(ϱ,u)(t)kL2(ϱ,u)(t)(k1)L2c1(1+t)34k2. (4.5)

    Therefore, we have completed the proof of this lemma.

    In order to obtain the decay rate of k+1nL2, employing operator to the equation (2.3)3, we have

    ntΔn=S3. (4.6)

    Then, we can deduce the lower bound decay rate of k+1n which is given in the following lemma.

    Lemma 4.2. Under all the assumptions of Theorem 1.3, then there is a positive constant c2 independent of time, such that for all tt and 0kN,

    k+1n(t)L2c2(1+t)34k2, (4.7)

    where t is a positive large time.

    Proof. If t is large enough, by using Lemma 3.2 and Lemma 7.5, for k=0,1, we have

    k+1n(t)L2k+1nl(t)L2c4(1+t)34k2Ct0(1+tτ)54k2S3L1dτc4(1+t)34k2Ct0(1+tτ)54k2(1+τ)2dτc4(1+t)34k2C(1+t)54k2c2(1+t)34k2. (4.8)

    Next, we employ interpolation inequality to establish the lower bound of decay rate for the higher–order spatial derivative of solution. The decay rates (4.8) together with the interpolations (k2)

    kfL2CfkL2f(k1)L2 (4.9)

    we have

    k+1n(t)L2c2(1+t)34k2, (4.10)

    for large time t. Thus the proof of Lemma 4.2 is completed.

    Therefore, we have completed the proof of Theorem 1.3.

    In this section, we will establish the decay rate for the mixed space–time derivatives of solution to the Cauchy problem (1.1)–(1.2).

    Lemma 5.1. Suppose that the assumptions of Theorem 1.1 and Theorem 1.2 are in force, the global solution (ϱ,u,n) hasthe time decay rate:

    ktu(t)L2C(1+t)54k2,k[0,N2],ltϱ(t)L2C(1+t)54l2,ltn(t)L2C(1+t)74l2,l[0,N1]. (5.1)

    Here, the positive constant C is independent of time.

    Proof. First of all, we shall estimate ktϱL2. For l=0,1,2,...,N1, taking l–th spatial derivative to (2.1)1, then multiplying the resulting identities by ltϱ and integrating over R3, we have

    ltϱ2L2=R3l(ϱdivu+uϱ+divu)ltϱdx:=K1+K2+K3. (5.2)

    We will estimate the right hand side of (5.2) term by term. First, for the term K1, by virtue of Lemma 7.3, and Young's inequality, we obtain

    K1(lϱL6uL3+ϱLl+1uL2)ltϱL2δltϱ2L2+Cl+1ϱ2L2u2L3+Cϱ2Ll+1u2L2δltϱ2L2+C(1+t)52l(1+t)3+C(1+t)3(1+t)52lδltϱ2L2+C(1+t)112l, (5.3)

    where δ is small enough. Similarly, we can get

    K2(luL6ϱL3+uLl+1ϱL2)ltϱL2δltϱ2L2+Cl+1u2L2ϱ2L3+Cu2Ll+1ϱ2L2δltϱ2L2+C(1+t)52l(1+t)3+C(1+t)3(1+t)52lδltϱ2L2+C(1+t)112l. (5.4)

    For the term K3, using Young's inequality, it is easy to see that

    K3δltϱ2L2+Cl+1u2L2δltϱ2L2+C(1+t)52l. (5.5)

    Combining (5.3), (5.4) with (5.5), we obtain

    ltϱ2L2(1+t)52l. (5.6)

    Then we shall estimate ktuL2. For k=0,1,2,...,N2, taking k–th spatial derivative to (2.1)2, multiplying the resulting identities by ktu and integrating over R3, we have

    ktu2L2=R3k[μΔu+(μ+λ)divuP(ρ)uuh(ϱ)[μΔu+(μ+ν)divu]f(ϱ)ϱg(ϱ)nΔn]ktudxδktu2L2+k[2uP(ρ)]2L2+k[uu+f(ϱ)ϱ]2L2+k[h(ϱ)2u]2L2+k[g(ϱ)nΔn]2L2=δktu2L2+L1+L2+L3+L4. (5.7)

    By applying Lemma 7.3 again, we estimate the second term in the right hand of (5.7).

    L1k+2u2L2+k+1ϱ2L2(1+t)72k+(1+t)52k(1+t)52k. (5.8)

    Similarly, for the term L2, we can bound

    L2k(ϱ,u)2L6(ϱ,u)2L3+k+1(ϱ,u)2L2(ϱ,u)2L(1+t)52k(1+t)3(1+t)112k. (5.9)

    For the term L3, it is easy to get

    L3kϱ2L62u2L3+k+2u2L2ϱ2L(1+t)52k(1+t)4+(1+t)72k(1+t)3(1+t)132k. (5.10)

    By employing the Leibniz formula and Lemma 7.3, we estimate the last term of (5.7)

    L4k[g(ϱ)nΔn]2L2g(ϱ)2Lk(nΔn)2L2+nΔn2L3kg(ϱ)2L6ϱ2Lk(nΔn)2L2+nΔn2L3k+1ϱ2L2ϱ2L(n2Lk+2n2L2+Δn2L3k+1n2L6)+n2L6Δn2L6k+1ϱ2L2(1+t)3(1+t)4(1+t)72k+(1+t)72(1+t)92(1+t)52k(1+t)52k. (5.11)

    Substituting estimates (5.8)–(5.11) into (5.7), we have

    ktu2L2(1+t)52k. (5.12)

    Similar to the estimate of the term ltϱL2. For l=0,1,2,...,N1, taking l–th spatial derivative to (2.1)3, then multiplying the resulting identities by ltn and integrating over R3, we have

    ltn2L2=R3l[Δnun+|n|2(n+ω0)]ltndxδltn2L2+l[Δn+un]2L2+l[|n|2(n+ω0)]2L2=δltn2L2+X1+X2. (5.13)

    By using Lemma 7.3 again, we estimate the second term of (5.13)

    X1l+2n2L2+lu2L6n2L3+l+1n2L2u2L(1+t)72l+(1+t)52l(1+t)3(1+t)72l. (5.14)

    By employing the Leibniz formula and Lemma 7.3, we estimate the last term of (5.13)

    X2=l[|n|2(n+ω0)]2L2l(|n|2)(n+ω0)2L2+lm=1Cmkm(n+ω0)lm(|n|2)2L2n2L3l+1n2L6+m(n+ω0)2L6lm(|n|2)2L3n2L3l+2n2L2+m+1n2L2n2L6l+1mn2L6(1+t)3(1+t)72l+(1+t)52m(1+t)72(1+t)72(lm)(1+t)72l. (5.15)

    Here, we have the basic fact that |n(x,t)+ω0|=|d(x,t)|=1, and

    l(|n|2(n+ω0))=l(|n|2)(n+ω0)+lm=1Cmkm(n+ω0)lm(|n|2). (5.16)

    Combining (5.14) with (5.15), we arrive at

    ltn2L2(1+t)72l. (5.17)

    Combining estimates (5.6), (5.12) with (5.17), then we complete the proof of this lemma.

    In this section, we will establish the lower bound of decay rate for the time derivative of solution of the system (2.1).

    Lemma 6.1. Under all the assumptions of Theorem 1.3, then the global solution (ϱ,u,n)has the time decay rate for all tt

    min{tϱ(t)L2,tu(t)L2,divu(t)L2}c3(1+t)54,tn(t)L2c3(1+t)54. (6.1)

    Here, t is a positive large constant, the positive constant c3 is independent of time.

    Proof. At first, we establish the lower bound time decay rate for tn in the L2–norm. With the help of the Eq (2.1)3, we have

    tnL2ΔnL2S3L2c2(1+t)54S3L2. (6.2)

    By applying the Sobolev inequality and decay rate (1.7), it is easy to get

    S3L2unL2+|n|2(n+ω0)L2CuLnL2+CnL3nL6CuH1nL2+CnH12nL2C(1+t)52. (6.3)

    And hence, we have

    tnL2c2(1+t)54C(1+t)52c3(1+t)54, (6.4)

    for all some large time t.

    Next, we establish lower bound time decay rate for tu in the L2–norm. Using the Eq (2.1)2, we can obtain

    ϱL2tuL2+2uL2+S2L2. (6.5)

    And hence, we have

    tuL2ϱL22uL2S2L2c1(1+t)54C(1+t)74S2L2. (6.6)

    By virtue of the Sobolev inequality and time decay rate (1.7), it is easy to deduce

    S2L2uuL2+h(ϱ)2uL2+f(ϱ)ϱL2+g(ϱ)nΔnL2C(ϱ,u)L(ϱ,u)L2+CϱL2uL2+CϱLnL32nL6C(ϱ,u)H1(ϱ,u)L2+CϱH12uL2+CϱH1nH13nL2C(1+t)52. (6.7)

    Together with estimates (6.6) and (6.7), yields directly

    tuL2c1(1+t)54C(1+t)74C(1+t)52c3(1+t)54. (6.8)

    Finally, we establish lower bound time decay rate for tϱ in the L2–norm. To accomplish this objective, we use the Eq (2.1)1 to obtain

    divuL2tϱL2+S1L2. (6.9)

    By virtue of the Sobolev inequality and decay rate (1.7), we can get

    S1L2ϱdivuL2+uϱL2C(ϱ,u)L(ϱ,u)L2C(ϱ,u)H1(ϱ,u)L2C(1+t)52, (6.10)

    and hence, we arrive at

    tϱL2divuL2C(1+t)52. (6.11)

    Now, we need to establish the lower bound decay rate for divuL2. Notice the relation differential relation Δ=div××, we get

    u2L2=divu2L2+×u2L2. (6.12)

    And hence, we have

    divuL2CuL2C×uL2c1C(1+t)54C×uL2, (6.13)

    which implies that we need to establish upper bound decay rate for ×uL2. To this end, we take the × operator the the velocity equation (2.1)2 to get

    t(×u)μΔ(×u)=×G2. (6.14)

    With the help of the Sobolev inequality, uniform bound (1.6), decay rate (1.7) and calculus identity (4.2), we have

    S2L1+S2L2CuuL1+Ch(ϱ)2uL1+Cf(ϱ)ϱL1+Cg(ϱ)nΔnL1+CuuL2+Ch(ϱ)2uL2+Cf(ϱ)ϱL2+Cg(ϱ)nΔnL2C((ϱ,u,n)L2(ϱ,u,n)H1+(ϱ,u,n)H1(ϱ,u,n)H2)Cδ0(1+t)54. (6.15)

    By virtue of the Duhamel principle formula, we obtain

    ×uL2C(1+t)54(Λ1F(×u0)L+Λ1F(×u0)L2)+Ct0(1+tτ)54(Λ1F(×S2)L+Λ1F(×S2)L2)dxC(δ0+δ1)(1+t)54+Cδ0t0(1+tτ)54(1+τ)54dxC(δ0+δ1)(1+t)54, (6.16)

    which, together with estimate (6.13) and smallness of δi(i=0,1), yields directly

    divuL2c1C(1+t)54C(δ0+δ1)(1+t)54c3(1+t)54. (6.17)

    This and the estimate (6.11) yields

    tϱL2c3(1+t)54C(1+t)52, (6.18)

    which implies directly

    tϱL2c3(1+t)54, (6.19)

    for some large time t. Therefore, we complete the proof of this lemma.

    For ease of use and clear reference, some Sobolev inequalities are listed as follows:

    Lemma 7.1. (i) If u(x)H1(R3), then the following inequalities hold:

    uL6Cu,uL3C(u+uL6)CuH1.

    (ii) Assume u(x)H2(R3), then

    uLCuH1.

    Proof. One can found them in [1].

    Now, we will introduce the Gagliardo-Nirenberg inequality that is frequently used in this paper.

    Lemma 7.2. Let 0i,jk; then we have

    ifLpjf1θLqkfθLr,

    where θ satisfies

    i31p=(j31q)(1θ)+(k31r)θ.

    Especially, while p=q=r=2, we have

    ifL2jfkikjL2kfijkjL2.

    Proof. The proof can be seen in [25].

    Next, to estimate the L2-norm of the spatial derivatives of the product of two functions, we shall use the following formula:

    Lemma 7.3. Let k1 be an integer; then one have

    k(fg)LpfLp1kgLp2+kfLp3gLp4,

    where p,p2,p3[1,+] and

    1p=1p1+1p2=1p3+1p4.

    Proof. For p=p2=p3=2, it can be proved by using Lemma 7.2. For the general case, one may to refer to [19].

    From Lemma 7.3, we can deduce the following commutator estimate:

    Lemma 7.4. Let f and g be smooth functions belonging to HkL for any integer k1 and define the commutator

    g=k(fg)fkg.

    Then we have

    [k,f]gLpfLp1k1gLp2+kfLp3gLp4.

    Here pi(i=1,2,3,4) are defined in Lemma 7.3.

    Finally, we introduce the estimate for the low–frequency part and the high–frequency part of f.

    Lemma 7.5. If 2p, fLp(R3), then we have

    flLp+fhLpfLp.

    Proof. For 2p, by Young's inequality for convolutions, for the low-frequency, we have

    flLpF1ϕL1fLpfLp,

    and hence

    fhLpfLp+flLpfLp.

    This work is partially supported by Guangxi Natural Science Foundation #2019JJG110003, #2019AC20214, and #2019JJA110071.

    The authors declare that they have no conflict of interest.



    [1] R. Adams, J. F. Fournier, Sobolev spaces, 2 Eds., New York: Academic Press, 2003.
    [2] S. J. Ding, J. L. Lin, C. Y. Wang, H. Y. Wen, Compressible hydrodynamic flow of liquid crystals in 1–D, Discrete Contin. Dyn. Syst., 32 (2012), 539–563. http://dx.doi.org/10.3934/dcds.2012.32.539 doi: 10.3934/dcds.2012.32.539
    [3] R. J. Duan, H. X. Liu, S. Ukai, T. Yang, Optimal LpLq convergence rate for the compressible Navier-Stokes quations with potential force, J. Differential Equations., 238 (2007), 220–223. https://doi.org/10.1016/j.jde.2007.03.008 doi: 10.1016/j.jde.2007.03.008
    [4] J. L. Ericksen, Hydrostatic theory of liquid crystals, Arch. Ration. Mech. Anal., 9 (1962), 371–378. https://doi.org/10.1007/BF00253358 doi: 10.1007/BF00253358
    [5] J. S. Fan, F. Jiang, Large-time behavior of liquid crystal flows with a trigonometric condition in two dimensions, Commun. Pure Appl. Anal., 15 (2016), 73–90. http://dx.doi.org/10.3934/cpaa.2016.15.73 doi: 10.3934/cpaa.2016.15.73
    [6] E. Feireisl, A. Novotný, H. Petzeltová, On the existence of globally defined weak solutions to the Navier-Stokes equations, J. Math. Fluid Mech., 3 (2001), 358–392. https://doi.org/10.1007/PL00000976 doi: 10.1007/PL00000976
    [7] Y. Guo, Y. J. Wang, Decay of dissipative equations and negative Sobolev spaces, Comm. Partial Differential Equations, 37 (2012), 2165–2208. https://doi.org/10.1080/03605302.2012.696296 doi: 10.1080/03605302.2012.696296
    [8] J. Gao, Q. Tao, Z. A. Yao, Long-time behavior of solution for the compressible nematic liquid crystal flows in R3, J. Differential Equations., 261 (2016), 2334–2383. https://doi.org/10.1016/j.jde.2016.04.033. doi: 10.1016/j.jde.2016.04.033
    [9] J. Gao, Z. Y. Lyu, Z. A. Yao, Lower bound of decay rate for higher–order derivatives of solution to the compressible fluid models of Korteweg type, Z. Angew. Math. Phys., 71 (2020), 1–19. https://doi.org/10.1007/s00033-020-01330-8 doi: 10.1007/s00033-020-01330-8
    [10] D. Hoff, Global solutions of the Navier–Stokes equations for multidimensional compressible flow with discontinuous initial data, J. Differential Equations., 200 (1995), 215–254. https://doi.org/10.1006/jdeq.1995.1111. doi: 10.1006/jdeq.1995.1111
    [11] X. P. Hu, H. Wu, Global solution to the three-dimensional compressible flow of liquid crystals, SIAM J. Math. Anal., 45 (2013), 2678–2699. https://doi.org/10.1137/120898814 doi: 10.1137/120898814
    [12] W. T. Huang, X. Y. Lin, W. W. Wang, Decay-in-time of the highest-order derivatives of solutions for the compressible isentropic MHD equations, J. Math. Anal. Appl., 502 (2021), 125273. https://doi.org/10.1016/j.jmaa.2021.125273. doi: 10.1016/j.jmaa.2021.125273
    [13] T. Huang, C. Y. Wang, H. Y. Wen, Blow up criterion for compressible nematic liquid crystal flows in dimension three, Arch. Ration. Mech. Anal., 204 (2012), 285–311. https://doi.org/10.1007/s00205-011-0476-1 doi: 10.1007/s00205-011-0476-1
    [14] X. D. Huang, J. Li, Z. P. Xin, Global well-posedness of classical solutions with large oscillations and vacuum to the three-dimensional isentropic compressible Navier–Stokes equations, Comm. Pure Appl. Math., 65 (2012), 549–585. https://doi.org/10.1002/cpa.21382 doi: 10.1002/cpa.21382
    [15] F. Jiang, S. Jiang, Asymptotic behaviors of global solutions to the two-dimensional non-resistive MHD equations with large initial perturbations, Adv. Math., 393 (2021), 108084. https://doi.org/10.1016/j.aim.2021.108084 doi: 10.1016/j.aim.2021.108084
    [16] F. Jiang, S. Jiang, D. H. Wang, On multi-dimensional compressible flows of nematic liquid crystals with large initial energy in a bounded domain, J. Funct. Anal., 265 (2013), 3369–3397. https://doi.org/10.1016/j.jfa.2013.07.026 doi: 10.1016/j.jfa.2013.07.026
    [17] F. Jiang, S. Jiang, D. H. Wang, Global weak solutions to the equations of compressible flow of nematic liquid crystals in two dimensions, Arch. Ration. Mech. Anal., 214 (2014), 403–451. https://doi.org/10.1007/s00205-014-0768-3 doi: 10.1007/s00205-014-0768-3
    [18] F. Jiang, G. C. Wu, X. Zhong, On exponential stability of gravity driven viscoelastic flows, J. Differential Equations., 260 (2016), 7498–7534. https://doi.org/10.1016/j.jde.2016.01.030. doi: 10.1016/j.jde.2016.01.030
    [19] N. Ju, Existence and uniqueness of the solution to the dissipative 2D quasi-geostrophic equations in the Sobolev space, Comm. Math. Phys., 251 (2004), 365–376. https://doi.org/10.1007/s00220-004-1062-2 doi: 10.1007/s00220-004-1062-2
    [20] F. M. Leslie, Some constitutive equations for liquid crystals, Arch. Ration. Mech. Anal., 28 (1968), 265–283. https://doi.org/10.1007/BF00251810 doi: 10.1007/BF00251810
    [21] H. Li, T. Zhang, Large time behavior of isentropic compressible Navier–Stokes system in R3, Math. Meth. Appl. Sci., 34 (2011), 670–682. https://doi.org/10.1002/mma.1391 doi: 10.1002/mma.1391
    [22] J. Li, Z. Xu, J. Zhang, Global existence of classical solutions with large oscillations and vacuum to the three–dimensional compressible nematic liquid crystal flows, J. Math. Fluid Mech., 20 (2018), 2105–2145. https://doi.org/10.1007/s00021-018-0400-7 doi: 10.1007/s00021-018-0400-7
    [23] A. Matsumura, T. Nishida, The initial value problems for the equations of motion of viscous and heat-conductive fluids, Proc. Jpn. Acad. Ser. A, 55 (1979), 337–342. https://doi.org/10.3792/pjaa.55.337 doi: 10.3792/pjaa.55.337
    [24] A. Matsumura, T. Nishida, The initial value problems for the equations of motion of viscous and heat-conductive gases, J. Math. Kyoto Univ., 20 (1980), 67–104. https://doi.org/10.1215/kjm/1250522322 doi: 10.1215/kjm/1250522322
    [25] L. Nirenberg, On elliptic partial diferential equations, Annali della Scuola Normale Superiore di Pisa, 13 (1959), 115–162. https://doi.org/10.1007/978-3-642-10926-3_1 doi: 10.1007/978-3-642-10926-3_1
    [26] M. E. Schonbek, Large time behaviour of solutions to the Navier–Stokes equations, Comm. Partial Differential Equations, 11 (1986), 733–763. https://doi.org/10.1080/03605308608820443 doi: 10.1080/03605308608820443
    [27] J. Wang, C. G. Xiao, Y. H. Zhang, Optimal large time behavior of the compressible Navier–Stokes–Korteweg system in R3, Appl. Math. Lett., 120 (2021), 107274. https://doi.org/10.1016/j.aml.2021.107274. doi: 10.1016/j.aml.2021.107274
    [28] Y. J. Wang, Z. Tan, Global existence and optimal decay rate for the strong solutions in H2 to the compressible Navier-Stokes equations, Appl. Math. Lett., 24 (2011), 1778–1784. https://doi.org/10.1016/j.aml.2011.04.028. doi: 10.1016/j.aml.2011.04.028
    [29] X. Y. Wang, W. W. Wang, On global behavior of weak solutions to the Navier–Stokes equations of compressible fluid for γ=5/3, Bound. Value Probl., 2015 (2015), 1–13. https://doi.org/10.1186/s13661-015-0443-0 doi: 10.1186/s13661-015-0443-0
    [30] W. W. Wang, On global behavior of weak solutions of compressible flows of nematic liquid crystals, Acta Math. Sci. Ser. B (Engl. Ed.), 35 (2015), 650–672. https://doi.org/10.1016/S0252-9602(15)30011-4 doi: 10.1016/S0252-9602(15)30011-4
    [31] W. W. Wang, Y. Y. Zhao, Time-decay solutions of the initial-boundary value problem of rotating magnetohydrodynamic fluids, Bound. Value Probl., 2017 (2017), 1–31. https://doi.org/10.1186/s13661-017-0845-2 doi: 10.1186/s13661-017-0845-2
    [32] G. C. Wu, Z. Tan, Global low–energy weak solution and large–time behavior for the compressible flow of liquid crystals, J. Differential Equations, 264 (2018), 6603–6632. https://doi.org/10.1016/j.jde.2018.01.045 doi: 10.1016/j.jde.2018.01.045
    [33] E. Zuazua, Time asymptotics for heat and dissipative wave equations, Preprint Boling Guo Institute of Applied Physics and Computational Mathematics PO Box, 8009, 2003. Available from: http://www.uam.es/enrique.zuazua.
  • 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(2069) PDF downloads(87) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog