Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in R3

  • Received: 01 June 2021 Revised: 01 July 2021 Published: 07 September 2021
  • 35B40, 76N10

  • We are concerned with the Cauchy problem of the 3D compressible Navier–Stokes–Poisson system. Compared to the previous related works, the main purpose of this paper is two–fold: First, we prove the optimal decay rates of the higher spatial derivatives of the solution. Second, we investigate the influences of the electric field on the qualitative behaviors of solution. More precisely, we show that the density and high frequency part of the momentum of the compressible Navier–Stokes–Poisson system have the same L2 decay rates as the compressible Navier–Stokes equation and heat equation, but the L2 decay rate of the momentum is slower due to the effect of the electric field.

    Citation: Guochun Wu, Han Wang, Yinghui Zhang. Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in R3[J]. Electronic Research Archive, 2021, 29(6): 3889-3908. doi: 10.3934/era.2021067

    Related Papers:

    [1] Guochun Wu, Han Wang, Yinghui Zhang . Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in $ \mathbb R^3 $. Electronic Research Archive, 2021, 29(6): 3889-3908. doi: 10.3934/era.2021067
    [2] Yuhui Chen, Ronghua Pan, Leilei Tong . The sharp time decay rate of the isentropic Navier-Stokes system in $ {\mathop{\mathbb R\kern 0pt}\nolimits}^3 $. Electronic Research Archive, 2021, 29(2): 1945-1967. doi: 10.3934/era.2020099
    [3] Yue Cao . Blow-up criterion for the 3D viscous polytropic fluids with degenerate viscosities. Electronic Research Archive, 2020, 28(1): 27-46. doi: 10.3934/era.2020003
    [4] Jingjing Zhang, Ting Zhang . Local well-posedness of perturbed Navier-Stokes system around Landau solutions. Electronic Research Archive, 2021, 29(4): 2719-2739. doi: 10.3934/era.2021010
    [5] Lirong Huang, Jianqing Chen . Existence and asymptotic behavior of bound states for a class of nonautonomous Schrödinger-Poisson system. Electronic Research Archive, 2020, 28(1): 383-404. doi: 10.3934/era.2020022
    [6] Xiu Ye, Shangyou Zhang . A stabilizer free WG method for the Stokes equations with order two superconvergence on polytopal mesh. Electronic Research Archive, 2021, 29(6): 3609-3627. doi: 10.3934/era.2021053
    [7] Jun Zhou . Initial boundary value problem for a inhomogeneous pseudo-parabolic equation. Electronic Research Archive, 2020, 28(1): 67-90. doi: 10.3934/era.2020005
    [8] Jiayi Han, Changchun Liu . Global existence for a two-species chemotaxis-Navier-Stokes system with $ p $-Laplacian. Electronic Research Archive, 2021, 29(5): 3509-3533. doi: 10.3934/era.2021050
    [9] Li Cai, Fubao Zhang . The Brezis-Nirenberg type double critical problem for a class of Schrödinger-Poisson equations. Electronic Research Archive, 2021, 29(3): 2475-2488. doi: 10.3934/era.2020125
    [10] Huafei Di, Yadong Shang, Jiali Yu . Existence and uniform decay estimates for the fourth order wave equation with nonlinear boundary damping and interior source. Electronic Research Archive, 2020, 28(1): 221-261. doi: 10.3934/era.2020015
  • We are concerned with the Cauchy problem of the 3D compressible Navier–Stokes–Poisson system. Compared to the previous related works, the main purpose of this paper is two–fold: First, we prove the optimal decay rates of the higher spatial derivatives of the solution. Second, we investigate the influences of the electric field on the qualitative behaviors of solution. More precisely, we show that the density and high frequency part of the momentum of the compressible Navier–Stokes–Poisson system have the same L2 decay rates as the compressible Navier–Stokes equation and heat equation, but the L2 decay rate of the momentum is slower due to the effect of the electric field.



    We consider the following 3D compressible Navier-Stokes- Poisson (NSP) system:

    {ρt+m=0,  xR3, t>0,mt+(mmρ)+p(ρ)+ρϕ=μΔ(mρ)+(μ+ν)((mρ)),λ2Δϕ=ρˉρ,  lim|x|ϕ(x,t)0,ρ(x,0)=ρ0(x), m(x,0)=m0(x),  xR3. (1.1)

    Here, the density ρ>0, m represents the momentum, u=mρ stands for the velocity, and ϕ represents the electrostatic potential. The viscosity coefficients μ>0 and ν satisfy 23μ+ν0, the Debye length λ>0 and the pressure functions p=p(ρ). ˉρ is the background doping profile, which is regarded as a positive constant for the sake of simplicity.

    To go directly to the theme of this paper, let us give some explanations about the above model. For the existence and long time behavior of the solution to the compressible Navier–Stokes (NS) system (i.e., ϕ0 in (1.1)), one can refer to [10,11,12,18,19,6] and references therein. For the compressible NSP system, Li–Matsumura–Zhang [14] proved that the density of the NSP system converges to its equilibrium state at the same L2–rate (1+t)34 as the compressible Navier–Stokes (NS) system, but the momentum of the NSP system decays at the L2–rate (1+t)14, which is slower than the L2–rate (1+t)34 for the compressible NS system due to the effect of the electric field. And then they extended similar result to the non–isentropic case [27]. Wang [25] obtained the optimal asymptotic decay of solutions just by pure energy estimates, and particularly he proved that the density of the compressible NSP system decays at the L2–rate (1+t)54, which is faster than the L2–rate (1+t)34 for the NS system due to the effect of the electric field. Hao–Li [9], Tan–Wu [22], Chikami–Danchin [4], Bie–Wang–Yao [2] and Shi-Xu [21] also established the unique global solvability and the optimal decay rates in critical spaces. We mention that there are many results on the existence and long time behavior of the weak solutions or non–constant stationary solutions, see, for example [1,5,8,28] and the references therein. For the compressible NSP system with external force, [16] investigated the existence and zero-electron-mass limit of strong solutions to the stationary with large external force and [17] proved that the strong solution existence to the boundary value problem in a bounded domain. For the bipolar Navier–Stokes–Poisson(BNSP) system, we refer to [7], [15], [24], [26], [29], [30], and references therein.

    We remark that this paper is strongly motivated by Li–Matsumura–Zhang [14]. Their main results can be outlined as follows: Assume that (ρ0ˉρ,m0)HlL1, l4, and (ρ0ˉρ,m0)HlL1 is sufficiently small. Then, the Cauchy problem (1.1) admits a global smooth solution (ρ,m,ϕ) satisfying:

    k(ρˉρ)(t)L2(1+t)34k2(ρ0ˉρ,m0)HlL1,fork=0,1, (1.2)
    k(m,ϕ)(t)L2(1+t)14k2(ρ0ˉρ,m0)HlL1,fork=0,1, (1.3)

    and

    (ρˉρ,m)(t)H4(ρ0ˉρ,m0)HlL1. (1.4)

    However, it is clear that (1.2)(1.4) give no information on the decay rates of higher order spatial derivatives (k2) of the solution. The main purpose of this article is to give a clear answer to this issue.

    In this paper, we first introduce some notations and conventions. We use Hl(R3) to denote the usual Sobolev spaces with norm Hl and 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 only dependent on the parameters of the problem. 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 by

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

    If the Fourier transform of f exists, then f=fh+fl.

    Now, we state out our main results in the following theorem:

    Theorem 1.1. Let p(ρ)>0 for ρ>0. Assume that (ρ0ˉρ,m0)H4(R3)L1(R3), with δ0=:(ρ0ˉρ,m0)H3(R3)L1(R3) small. Then for any t0, the Cauchy problem (1.1) has a unique global solution (ρ,m,ϕ) satisfying the following optimal time decay rates

    k(ρˉρ)L2(1+t)34k2, (1.5)
    k(m,ϕ)L2(1+t)14k2, (1.6)

    and

    k(mh,ϕh)L2(1+t)34k2, (1.7)

    for 2k4.

    Remark 1.2. Compared to the result (1.2)-(1.4) in [14], the L2 decay rates of higher spatial derivatives from 2-order to 4-order of the solution (ρ,m,ϕ) in (1.5)-(1.7) are totally new. Furthermore, we only need the smallness of H3L1-norm of the initial data, while H4-norm of the initial data may be arbitrarily large. Finally, by noticing that, for 2k4, we can obtain the optimal decay rates of k(mh,ϕh)L2 in (1.7), which are the same as those of the compressible Navier–Stokes equations and heat equation, and particularly faster than one of k(m,ϕ)L2 in (1.6). This implies that the electric field has no effect on the decay rates of the density and the high-frequency part of the momentum, but reduces the decay rate of the momentum.

    Now, let us sketch the strategy of proving Theorem 1.1. Motivated by the results in (1.2)-(1.4), we will focus on deriving the L2 time decay rates of higher spatial derivatives from 2-order to 4-order of the solution (ρ,m,ϕ). Our strategy can be outlined as follows. First, when we prove the Theorem 1.1, we encounter such a difficulty: we need the optimal decay rate of the second-order spatial derivative of the solution to deduce the decay rate of the third-order spatial derivative of the solution. In the same way, we also need the decay rate of the third-order spatial derivative of the solution to derive the decay rate of the fourth-order spatial derivative of the solution. Therefore, we will prove Theorem 1.1 through three cases. Notice that the optimal decay rates on low-frequency part of (ρ,m,ϕ) has been established by [3]. Therefore, we only need to derive the optimal decay rates on k(ρh,mh,ϕh), k=2,3,4. For case 1, we hope to establish the L2-time decay rates of second-order spatial derivative of (ρ,m,ϕ). We rewrite system (1.1) into (2.1), and deduce the energy inequality as follows:

    12ddtR3|2nh|2+|2uh|2+|2ϕh|2dx+(μ1+μ2)R3|3uh|2dxδ0(2n2L2+2ϕh2L2+3u2L2. (1.8)

    However, it should be mentioned that in the process of deducing the energy inequality (1.8), we encounter the trouble term I2=2(nu)h,2nh in (2.7). If we estimate it directly, we will meet the norm 3nL2 which however can not be controlled in our framework. To overcome this obstacle, we need to reduce the order of the density by splitting I2 into three parts as follows:

    2(nu)h,2nh=2(nu)2(nu)l,2nh=2(nhu)+2(nlu)2(nu)l,2nh (1.9)
    :=I21+I22+I23,

    and then make full use of the benefit of the low–frequency and high–frequency decomposition technique to bound the terms I21, I22 and I23 one by one. For I21, it is clear that

    2(nhu)=u3nh+2u2nh+2unh,

    and using integration by parts, we have

    u3nh,2nh=12R3divu|2nh|2dx.

    Thus we can deduce

    I2δ0(2n2L2+2nh2L2+3u2L2) (1.10)

    as in (2.13). Now, note that (1.8) only gives the dissipation estimate for uh. In order to explore the dissipation estimates for nh and ϕh, we will employ the new interactive energy functional between uh and ρh by using critical L2 estimates, low-frequency and high-frequency decomposition. Particularly, we can get

    ddtR3uh2nhdx+R3|2nh|2|2uh|2+|2ϕh|2dxδ0(2n2L2+3u2L2+2uh2L2+2nh2L2)+(2nh2L2+3uh2L2). (1.11)

    Next, we choose a sufficiently large positive constant D1 and then define the temporal energy functional as

    E1(t)=D12(2nh2L2+2uh2L2+2Eh2L2)+R3uh2nhdx. (1.12)

    Noting that (1.12) is equivalent to 2nh2L2+2uh2L2+2Eh2L2. Multiplying (1.8) by D1, adding the resulting inequality with (1.11), we obtain

    ddtE1(t)+C1E1(t)2nl2L2+3ul2L2, (1.13)

    then the optimal L2 time decay rates of 2((ρˉρ)h,mh,ϕh)L2 in (1.7) for k=2 can be deduced by virtue of Lemma 2.1, Lemma 2.2, Gronwall's argument. Furthermore, combining with the 2-order low-frequency decay estimates in Lemma 2.2, we get the decay rates of 2-order spatial derivatives of (ρ,m,ϕ) immediately. This implies that we complete the proof of Theorem 1.1 for k=2. We use similar methods to prove the Theorem 1.1 for k=3 and k=4 in case 2 and case 3 respectively.

    This section is devoted to prove the optimal L2 time decay rates of solution stated in Theorem 1.1. First, we rewrite the system. Denoting (ρ,u)=(1+n,mρ) and ϕ=E, then the Cauchy problem (1.1) can be reformulated as

    {nt+u=f1,ut+n+Eμ1Δuμ2(u)=f2f3,E=(Δ)1n,  lim|x|E0,n(x,0)=n0(x)=:ρ0(x)1,  u(x,0)=u0(x)=m0(x)/ρ0(x),  xR3, (2.1)

    where

    f1=nunu,
    f2=(u)u+(1p(1+n)1+n)n+μ1(n1+n)Δu+μ2(n1+n)(u),
    f3=nϕ=nE,

    and as in [14], we have taken ˉρ=1, p(1)=1, μ=μ1, (μ+ν)=μ2 and d=1 for simplicity. Notice that

    f2O(1)(nn+uu+n2u),

    and

    kEL2k1nL2, k1.

    Second, the decay rates on the linearized system of (2.1) are given in the following lemma:

    Lemma 2.1 Assume that U0=(n0,m0)Hl(R3)L1(R3), l4, and denote (ˉn(t),ˉm(t))=:ˉU(t). Then, for 0kl, the solution (ˉn,ˉm,ˉE) of the linearized system of (1.1) with ˉE=(Δ)1ˉn satisfies that

    kˉn(t)L2(1+t)34k2(U0L1+kU0L2), (2.2)
    k(ˉE,ˉm)(t)L2(1+t)14k2(U0L1+kU0L2). (2.3)

    Proof. See [14].

    Third, we state the L2-time decay estimates on the low-frequency part of the solution in the nonliear system (1.1).

    Lemma 2.2. Assume that the assumptions of Theorem 1.1 are in force, Then for 0kl, the following decay rates hold:

    knl(t)L2(1+t)3+2k4, (2.4)
    k(El,ml)(t)L2(1+t)1+2k4. (2.5)

    Proof. See [3].

    Next, we will prove Theorem 1.1 in three cases. Before giving the precise proofs of these three cases, we must emphasize that under the assumption of Theorem 1.1, [14] proved that the following estimate holds

    (ρˉρ,m)(t)H4δ0. (2.6)

    More details of (2.6), please refer to (3.62)-(3.64) on page 697 of reference [14].

    Case 1. Proof of Theorem 1.1 for k=2. Theorem 1.1 will be proved by the good properties of the low-frequency and high-frequency decomposition. The proof involves the following steps.

    Step 1. High–frequency L2 energy estimate. Taking

    F1(1ϕ(ξ))2(2.1)1,2nh+F1(1ϕ(ξ))2(2.1)2,2uh,

    and using integration by parts, yields directly

    12ddtR3|2nh|2+|2uh|2+|2Eh|2dx+(μ1+μ2)R3|3uh|2dx=2(nu)h,2nh2(nu)h,2nhEh,2(nu)hEh,2(nu)h2(f2+f3)h,2uh:=5i=1Ii. (2.7)

    The right-hand side of (2.7) can be estimated one by one. For term I1, due to Hölder's inequality, Young inequality, Sobolev interpolation theorem, it holds that

    |I1|=|2(nu)h,2nh|2(nu)L22nhL2(uL2nL2+nL3uL2)2nhL2(uH22nL2+nH23uL2)2nhL2δ0(2n2L2+3u2L2). (2.8)

    The second term I2 can be rewritten as follows

    I2=2(nu)h,2nh=2(nu)2(nu)l,2nh=2(nhu)+2(nlu)2(nu)l,2nh:=I21+I22+I23. (2.9)

    It is easy to obtain

    2(nhu)=u3nh+2u2nh+2unh.

    Following from integration by parts, we have

    u3nh,2nh=12R3divu|2nh|2dx,

    thus

    |I21|(uL2nhL2+nhL32uL6)2nhL2(uH22nhL2+nH13uL2)2nhL2δ0(2n2L2+3u2L2). (2.10)

    For the term I22, we have

    |I22|=|2(nlu),2nh|2(nlu)L22nhL2(uL3nlL2+nlL32uL6)2nhL2(uH22nL2+nH13uL2)2nhL2δ0(2n2L2+3u2L2+2nh2L2). (2.11)

    For the term I23, by Lemma A.4, Hölder's inequality, Young inequality, we have

    |I23|=|2(nu)l,2nh|2(nu)lL22nhL2nuL22nhL2uL3nL62nhL2uH12nL22nhL2δ0(2n2L2+2nh2L2), (2.12)

    Substituting (2.10)-(2.12) into (2.9), we can conclude that

    I2δ0(2n2L2+2nh2L2+3u2L2) (2.13)

    For the term I3 and I4, similar to the proofs of (2.8) and (2.13), it holds that

    |I3|+|I4|=|Eh,2(nu)h|+|Eh,2(nu)h|2EhL2(2(nu)L2+2(nu)hL2)δ0(2Eh2L2+2n2L2+3u2L2), (2.14)

    For the term I5, making use of integration by parts, we obtain

    I5=2(f2+f3)h,2uh=(f2+f3)h,3uh=(nn)h+(uu)h+(n2u)h(nE)h,3uh:=I51+I52+I53+I54. (2.15)

    For the term I51, we have

    |I51|=|(nn)h,3uh|(nn)L23uhL2(nL3nL6+nL2nL2)3uhL2(nH12nL2+nH22nL2)3uhL2δ0(2n2L2+3uh2L2). (2.16)

    For the term I52, it holds that

    |I52|=|(uu)h,3uh|2(uu)L23uhL2(uL3uL2+uL32uL6)3uhL2(uH23uL2+uH13uL2)3uhL2δ0(3u2L2+3uh2L2). (2.17)

    For the term I53, we get

    |I53|=|(n2u)h,3uh|(n2u)L23uhL2(nL3uL2+2uL3nL6)3uhL2(nH23uL2+2uH12nL2)3uhL2δ0(3u2L2+2n2L2). (2.18)

    For the term I54, we have

    |I54|=|(nE)h,3uh|2(nE)L23uhL2(EL2nL2+nL32EL6)3uhL2(EH12nL2+nH12nL2)3uhL2(nH12nL2+nH12nL2)3uhL2δ0(2n2L2+3uh2L2). (2.19)

    Thus, one may deduce that

    |I5|δ0(2n2L2+3u2L2). (2.20)

    Substituting (2.8), (2.13)–(2.14) and (2.20) into (2.7) and using the smallness of δ0, yield directly

    12ddtR3|2nh|2+|2uh|2+|2Eh|2dx+(μ1+μ2)R3|3uh|2dxδ0(2n2L2+2Eh2L2+3u2L2). (2.21)

    Step 2. Dissipation of 2nh. Taking F1(1ϕ(ξ))(2.1)2,2nh, it holds that

    ddtR3uh2nhdx+R3|2nh|2|2uh|2+|2Eh|2dx=2(nu)h,uh2(nu)h,uh(f2+f3)h2nh+(μ1+μ2)3uh,2nh:=I6+I7+I8+I9. (2.22)

    We should estimate the right-hand side of (2.22) one by one. For I6, using a similar argument like (2.8), we have

    |I6|δ0(2n2L2+3u2L2+2uh2L2). (2.23)

    For I7, similar to the proof of (2.13), it holds that

    |I7|δ0(2n2L2+2uh2L23u2L2+2nh2L2). (2.24)

    For I8, as the proof of (2.19), we have

    |I8|δ0(2n2L2+3u2L2. (2.25)

    For I9, we obtain

    |I9|(μ1+μ2)3uhL22nhL2142nh2L2+3uh2L2. (2.26)

    Putting (2.23)–(2.26)into (2.22) and using the smallness of δ0, we conclude that

    ddtR3uh2nhdx+R3|2nh|2+|2Eh|2dxδ0(2n2L2+3u2L2+2uh2L2+2nh2L2)+142nh2L2+3uh2L2+R3|2uh|2dx. (2.27)

    Step 3. Closing the estimates. Now, we are in a position to close the estimates. To do this, we choose sufficiently large time T1 and positive constant D1, and then define the temporary energy functional

    E1(t)=D12(2nh2L2+2uh2L2+2Eh2L2)+R3uh2nhdx, (2.28)

    for tT1, it is noticed that E1(t) is equivalent to 2nh2L2+2uh2L2+2Eh2L2 since D1 is large enough. Substituting (2.21) and (2.27) into

    D1×(2.21)+(2.27),

    which together with the smallness of δ0, for tT1, it holds that

    ddtE1(t)+(34δ0)2nh2L2+[(μ1+μ2)D12]3uh2L2+2Eh2L22nl2L2+3ul2L2, (2.29)

    where we have used the fact that T1 is large enough. On other hand, it is clear that

    (34δ0)2nh2L2+[(μ1+μ2)D12]3uh2L2+2Eh2L2C1E1(t). (2.30)

    Hence, by virtue of (2.4)-(2.5), (2.29)-(2.30) and Gronwall's inequality, we can arrive at

    2nh2L2+2uh2L2+2Eh2L2(1+t)74. (2.31)

    Furthermore, combining with (2.4)-(2.5) and (2.31), the following estimates can be obtained:

    2n2L22nh2L2+2nl2L2(1+t)74, (2.32)

    and

    2(u,E)2L22(uh,Eh)2L2+2(ul,El)2L2(1+t)54. (2.33)

    The proof of Theorem 1.1 for k=2 is completed.

    Case 2. Proof of Theorem 1.1 for k=3. Theorem 1.1 for k=3 will be proved as in Case 1 by the good properties of the low-frequency and high-frequency decomposition. The proof involves the following steps.

    Step 1. High–frequency L2 energy estimate. Taking

    F1(1ϕ(ξ))3(2.1)1,3nh+F1(1ϕ(ξ))3(2.1)2,3uh,

    and using integration by parts, which implies

    12ddtR3|3nh|2+|3uh|2+|3Eh|2dx+(μ1+μ2)R3|4uh|2dx=3(nu)h,3nh3(nu)h,3nh2Eh,3(nu)h2Eh,3(nu)h3(f2+f3)h,3uh:=5i=1Ji. (2.34)

    The five terms in the above equation can be estimated as follows. Firstly, for term J1, using Hölder's inequality, Young inequality, Sobolev interpolation theorem, we obtain

    |J1|=|3(nu)h,3nh|3(nu)L23nhL2(uL3nL2+nL4uL2)3nhL2(uH23nL2+nH24uL2)3nhL2δ0(3n2L2+3nh2L2+4u2L2), (2.35)

    The term J2 can be rewritten as follows

    J2=3(nu)h,3nh=3(nu)3(nu)l,3nh=3(nhu)+3(nlu)3(nu)l,3nh:=J21+J22+J23. (2.36)

    It is obvious that

    3(nhu)=u4nh+3u3nh+32u2nh+nh3u.

    By virtue of integration by parts, we arrive at

    u4nh,3nh=12R3divu|3nh|2dx,

    thus

    |J21|(uL3nhL2+2uL32nhL6+nhL33uL6)3nhL2(uH23nhL2+2uH13nhL2+nH14uL2)3nhL2δ0(3nh2L2+4u2L2), (2.37)

    For the term J22, we have

    |J22|=|3(nlu),3nh|3(nlu)L23nhL2(uL4nlL2+nlL33uL6)3nhL2(uH23nL2+nH14uL2)3nL2δ0(3n2L2+4u2L2). (2.38)

    For the term J23, with the help of LemmaA.1-Lemma A.3, Hölder's inequality, (1.3), (2.32)-(2.33) and Young inequality, we have

    |J23|=|3(nu)l,3nh|(nu)L23nhL2(uL32nL6+uL3nL6)3nhL2uH13nL23nhL2+u14L22u34L22nL23nhL2δ0(3n2L2+3nh2L2)+(1+t)14×14(1+t)54×34(1+t)743nhL2δ0(3n2L2+3nh2L2)+(1+t)94123nhL2δ0(3n2L2+3nh2L2)+(1+t)92+(1+t)13nh2L2. (2.39)

    Substituting (2.37)-(2.39) into (2.36), we can achieve

    J2δ0(3n2L2+3nh2L2+4u2L2)+(1+t)92+(1+t)13nh2L2. (2.40)

    For the term J3 and J4, similarly to the proof of (2.35) and (2.38), it holds that

    |J3|+|J4|δ0(3Eh2L2+3n2L2+4u2L2)+(1+t)92+(1+t)13Eh2L2, (2.41)

    For the term J5, by integration by parts, we have

    J5=3(f2+f3)h,3uh=2(f2+f3)h,4uh=2(nn)h+(uu)h+(n2u)h(nE)h,4uh:=J51+J52+J53+J54. (2.42)

    For the term J51, we have

    |J51|=|2(nn)h,4uh|2(nn)L24uhL2(nL32nL6+nL3nL2)4uhL2(nH13nL2+nH23nL2)4uhL2δ0(3n2L2+4uh2L2). (2.43)

    For the term J52, we have

    |J52|=|2(uu)h,4uh|3(uu)L24uhL2(uL4uL2+uL33uL6)4uhL2(uH24uL2+uH14uL2)4uhL2δ0(4u2L2+4uh2L2). (2.44)

    For the term J53, we have

    |J53|=|2(n2u)h,4uh|2(n2u)L24uhL2(nL4uL2+2nL62uL3)4uhL2(nH24uL2+3nL22uH1)4uhL2δ0(4u2L2+3n2L2). (2.45)

    For the term J54, we have

    |J54|=|2(nE)h,4uh|3(nE)L24uhL2(EL3nL2+nL33EL6)4uhL2(EH13nL2+nH13nL2)4uhL2(nH13nL2+nH13nL2)4uhL2δ0(3n2L2+4uh2L2). (2.46)

    Thus, we can immediately to obtain

    |J5|δ0(3n2L2+4u2L2). (2.47)

    Substitute (2.35), (2.40)–(2.41) and (2.47) into (2.34) and use the smallness of δ0, we deduce that in fact

    12ddtR3|3nh|2+|3uh|2+|3Eh|2dx+(μ1+μ2)R3|4uh|2dxδ0(3Eh2L2+4u2L2+3n2L2)+(1+t)92+(1+t)13nh2L2+(1+t)13Eh2L2. (2.48)

    Step 2. Dissipation of 3nh. Applying the operator 2F1(1ϕ(ξ)) to (2.1)2, multiplying the resulting equality by 3nh, integrating over R3, it holds that

    ddtR32uh3nhdx+R3|3nh|2|3uh|2+|3Eh|2dx=3(nu)h,uh3(nu)h,uh2(f2+f3)h3nh+(μ1+μ2)4uh,3nh:=J6+J7+J8+J9. (2.49)

    For the term J6, we have

    |J6|δ0(3n2L2+4u2L2+3uh2L2). (2.50)

    For the term J7, similarly to the proof of (2.40), we obtain

    |J7|δ0(3n2L2+4u2L2)+(1+t)92+(1+t)13uh2L2. (2.51)

    For the term J8, similarly to the proof of (2.42), it holds that

    |J8|δ0(3n2L2+4uh2L2+3uh2L2). (2.52)

    For J9, we obtain

    |J9|(μ1+μ2)4uhL23nhL2143nh2L2+4uh2L2. (2.53)

    Substituting (2.50)–(2.53)into (2.49) and making use of the smallness of δ0, we conclude that

    ddtR32uh3nhdx+R3|3nh|2+|3Eh|2dxδ0(3n2L2+3nh2L2+3uh2L2+4u2L2)+(1+t)92+(1+t)13uh2L2+143nh2L2+4uh2L2+R3|3uh|2dx. (2.54)

    Step 3. Closing the estimates. Now, let us close the estimates. For any t0, we define the temporary energy functional as follows

    E2(t)=D22(3nh2L2+3uh2L2+3Eh2L2)+R32uh3nhdx, (2.55)

    for tT2, where it is noticed that E2(t) is equivalent to 3nh2L2+3uh2L2+3Eh2L2 if we choose sufficiently large time T2 and positive constant D2. Putting (2.48) and (2.54) into

    D2×(2.48)+(2.54),

    for tT2, δ0 is small enough implies

    ddtE2(t)+(34δ0)3nh2L2+[(μ1+μ2)D22]4uh2L2+3Eh2L23nl2L2+4ul2L2, (2.56)

    where we have used the fact that T2 is large enough. On other hand, it is clear that

    (34δ0)3nh2L2+[(μ1+μ2)D22]4uh2L2+3Eh2L2C2E2(t). (2.57)

    Thus, in view of (2.4)-(2.5), (2.56)-(2.57) and Gronwall's argument, we have

    3nh2L2+3uh2L2+3Eh2L2(1+t)94. (2.58)

    Furthermore, utilizing (2.4)-(2.5) and (2.58), it is easy to have

    3n2L23nh2L2+3nl2L2(1+t)94, (2.59)
    3(u,E)2L23(uh,Eh)2L2+3(ul,Eh)2L2(1+t)74. (2.60)

    This complete the proof of Theorem 1.1 for k=3.

    Case 3. Proof of Theorem 1.1 for k=4. taking the similar argument as in Case 2, we can prove Theorem 1.1 for k=4 by the following steps.

    Step 1. High–frequency L2 energy estimate. Taking

    F1(1ϕ(ξ))4(2.1)1,4nh+F1(1ϕ(ξ))4(2.1)2,4uh,

    and applying integration by parts, we have

    12ddtR3|4nh|2+|4uh|2+|4Eh|2dx+(μ1+μ2)R3|5uh|2dx=4(nu)h,4nh4(nu)h,4nh3Eh,4(nu)h3Eh,4(nu)h4(f2+f3)h,4uh:=5i=1Ki. (2.61)

    We will estimate the right-hand side of the above equation term by term. First, for term K1, from Hölder's inequality, Young inequality, Sobolev interpolation theorem, it leads to

    |K1|=|4(nu)h,4nh|4(nu)L24nhL2(uL4nL2+nL5uL2)4nhL2(uH24nL2+nH25uL2)4nhL2δ0(4n2L2+4nh2L2+5u2L2), (2.62)

    For the term K2, we can rewrite it as

    K2=4(nu)h,4nh=4(nu)4(nu)l,4nh=4(nhu)+4(nlu)4(nu)l,4nh:=K21+K22+K23. (2.63)

    Notice that

    4(nhu)=u5nh+4u4nh+62u3nh+43u2nh+4unh.

    From integration by parts, one has

    u5nh,3nh=12R3divu|4nh|2dx.

    Note that

    3u2nh,4nh=3uh2nh,4nh+3ul2nh,4nh,

    and

    4unh,4nh=4uhnh,4nh+4ulnh,4nh.

    Thus, we can have the following estimate

    |K21|(uL4nhL2+2uL33nhL6+2nhL33uhL6+2nhL33ulL6+nhL34uhL6+4ulL3nhL6)4nhL2(uH24nhL2+2uH14nhL2+2nH14uhL2+2nH14uL2+nH15uhL2+uH14nhL2)4nhL2δ0(4n2L2+5u2L2), (2.64)

    For the term K22, we have

    |K22|=|4(nlu),4nh|4(nlu)L24nhL2(uL5nlL2+nlL34uL6)4nhL2(uH24nL2+nH15uL2)4nhL2δ0(4n2L2+5u2L2+4nh2L2). (2.65)

    For the term K23, we achieve

    |K23|=|4(nu)l,4nh|4(nu)lL24nhL22(nu)L24nhL2(uL33nL6+nL62uL3)4nhL2(uH14nL2+2nL2u14L23u34L2)4nhL2δ0(4n2L2+4nh2L2)+(1+t)114124nhL2δ0(4n2L2+4nh2L2)+(1+t)112+(1+t)14nh2L2. (2.66)

    Substituting (2.64)-(2.66) into (2.63), we can arrive at

    K2δ0(4n2L2+4nh2L2+5u2L2)+(1+t)112+(1+t)14nh2L2. (2.67)

    For the term K3 and K4, similarly to the proof of (2.62) and (2.65), it holds that

    |K3|+|K4|δ0(4Eh2L2+4n2L2+5u2L2)+(1+t)112+(1+t)14Eh2L2, (2.68)

    For the term K5, we have

    K5=3(f2+f3)h,5uh=3(f2+f3)h,5uh=3(nn)h+(uu)h+(n3u)h(nE)h,5uh:=K51+K52+K53+K54. (2.69)

    For the term K51, we have

    |K51|=|3(nn)h,5uh|3(nn)L25uhL2(nL33nL6+nL4nL2)5uhL2(nH14nL2+nH24nL2)5uhL2δ0(4n2L2+5uh2L2). (2.70)

    For the term K52, we have

    |K52|=|3(uu)h,5uh|4(uu)L25uhL2(uL5uL2+uL34uL6)5uhL2(uH25uL2+uH15uL2)5uhL2δ0(5u2L2+5uh2L2). (2.71)

    For the term K53, we obtain

    |K53|=|3(n2u)h,5uh|3(n2u)L25uhL2(nL5uL2+2uL33nL6)5uhL2(nH25uL2+2uH14nL2)5uhL2δ0(5u2L2+4n2L2). (2.72)

    For the term K54, we have

    |K54|=|3(nE)h,5uh|4(nE)L25uhL2(EL4nL2+nL34EL6)5uhL2(EH14nL2+nH14nL2)5uhL2(nH14nL2+nH14nL2)5uhL2δ04n2L2+5uh2L2. (2.73)

    Thus, we can arrive at

    |K5|δ0(4n2L2+5u2L2). (2.74)

    Substituting (2.62), (2.67)–(2.68) and (2.74) into (2.61) and using the smallness of δ0, we conclude that

    12ddtR3|4nh|2+|4uh|2+|4Eh|2dx+(μ1+μ2)R3|5uh|2dxδ0(4Eh2L2+5u2L2+4n2L2)+(1+t)92+(1+t)14nh2L2+(1+t)14Eh2L2. (2.75)

    Step 2. Dissipation of 4nh. Applying the operator 3F1(1ϕ(ξ)) to (2.1)2, multiplying the resulting equality by 4nh, integrating over R3, we have

    ddtR33uh4nhdx+R3|4nh|2|4uh|2+|4Eh|2dx=4(nu)h,3uh4(nu)h,3uh4(f2+f3)h,4nh+(μ1+μ2)5uh,4nh:=K6+K7+K8+K9. (2.76)

    For the term , we have

    (2.77)

    For the term , similarly to the proof of (2.67), we obtain

    (2.78)

    For the term , similarly to the proof of (2.69), it holds that

    (2.79)

    For , we have

    (2.80)

    Substituting (2.77)–(2.80) into (2.76) and using the smallness of , we dedeuce that

    (2.81)

    Step 3. Closing the estimates. Now, we are in a position to close the estimates. To do this, we define the temporal energy functional

    (2.82)

    for , where it is noticed that is equivalent to if we choose sufficiently large time and positive constant . Substituting (2.75) and (2.81) into

    and using the smallness of , for , we have

    (2.83)

    since is large enough. On other hand, it is clear that

    (2.84)

    Therefore, together with (2.4)-(2.5), (2.83)-(2.84) and Gronwall's argument, we have

    (2.85)

    Consequently, we have from (2.4)-(2.5) and (2.85) that

    (2.86)
    (2.87)

    This completes the proof of Theorem 1.1 for .

    Therefore, we complete the proof of Theorem 1.1.

    Appendix A. Analytic tools. We will use the Sobolev interpolation of Gagliardo-Nirenberg inequality.

    Lemma A.1. Let ; then we have

    where satisfies

    Lemma A.2. Let be an integer; then one has

    where and

    Finally, we introduce the lemma concerning the estimate for the low-frequency part and the high-frequency part of .

    Lemma A.3. If , , then we have

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

    and hence

    Guochun Wu's research was in part supported by National Natural Science Foundation of China (No. 11701193, 11671086), Natural Science Foundation of Fujian Province (No. 2018J05005, 2017J01562), Program for Innovative Research Team in Science and Technology in Fujian Province University Quanzhou High–Level Talents Support Plan (No. 2017ZT012). Yinghui Zhang' research is partially supported by Guangxi Natural Science Foundation (No. 2019JJG110003, 2019AC20214), and National Natural Science Foundation of China (No. 11771150, 11571280, 11301172 and 11226170.)



    [1] Long time behavior of weak solutions to Navier–Stokes–Poisson system. J. Math. Fluid Mech. (2012) 14: 279-294.
    [2] Optimal decay rate for the compressible Navier–Stokes–Poisson system in the critical framework. J. Differential Equations (2017) 263: 8391-8417.
    [3] Q. Chen, G. Wu and Y. Zhang, Optimal large time behavior of the compressible bipolar Navier-Stokes-Poisson system with unequal viscosities, Preprint, arXiv: 2104.08565v1 [math.AP] 17 Apr 2021.
    [4] On the global existence and time decay estimates in critical spaces for the Navier–Stokes–Poisson system. Math. Nachr. (2017) 290: 1939-1970.
    [5] Local and global existence for the coupled Navier–Stokes–Poisson problem. Quart. Appl. Math. (2003) 61: 345-361.
    [6] R. Duan, et al. Optimal - convergence rates for the compressible Navier-Stokes equations with potential force, J. Differential Equations, 238 (2007), 220-33. doi: 10.1016/j.jde.2007.03.008
    [7] Stability of rarefaction wave and boundary layer for outflow problem on the two–fluid Navier–Stokes–Poisson equations. Commun. Pure Appl. Anal. (2013) 12: 985-1014.
    [8] Global in time weak solution for compressible barotropic self-gravitating fluids. Discrete Contin. Dyn. Syst. (2004) 11: 113-130.
    [9] Global existence for compressible Navier-Stokes-Poisson equations in three and higher dimensions. J. Differential Equations (2009) 246: 4791-4812.
    [10] On large time behavior of solutions to the compressible Navier-Stokes equations in the half space in . Arch. Rational Mech. Anal. (2002) 165: 89-159.
    [11] Asymptotic behavior of solutions of the compressible Navier-Stokes equations on the half space. Arch. Ration. Mech. Anal. (2005) 177: 231-330.
    [12] Some estimates of solutions for the equations of motion of compressible viscous fluid in an exterior domain in . J. Differ. Equ. (2002) 184: 587-619.
    [13] Decay estimates of solutions for the equations of motion of compressible viscous and heat-conductive gases in an exterior domain in . Comm. Math. Phys. (1999) 200: 621-659.
    [14] Optimal decay rate of the compressible Navier-Stokes-Poisson system in . Arch. Ration. Meth. Anal. (2010) 196: 681-713.
    [15] Time asymptotic behavior of the bipolar Navier–Stokes–Poisson system. Acta Math. Sci. B (2009) 29: 1721-1736.
    [16] Y. Li and J. Liao, Existence and zero-electron-mass limit of strong solutions to the stationary compressible Navier-Stokes-Poisson equation with large external force, Math. Methods Appl. Sci., 41 (2018), 646–663. doi: 10.1002/mma.4634
    [17] Y. Li and N. Zhang, Decay rate of strong solutions to compressible Navier-Stokes-Poisson equations with external force, Electron. J. Differential Equations, (2019), Paper No. 61, 18 pp.
    [18] The initial value problem for the equations of motion of compressible viscous and heat conductive fluids. Proc. Japan Acad. Ser. A Math. Sci. (1979) 55: 337-342.
    [19] The initial value problem for the equations of motion of viscous and heat–conductive gases. J. Math. Kyoto Univ. (1980) 20: 67-104.
    [20] A. Matsumura and T. Nishida, Initial boundary value problems for the equations of motion of compressible viscous and heat conductive fluids, Comm. Math. Phys., 89 (1983), 445–464. doi: 10.1007/BF01214738
    [21] A sharp time–weighted inequality for the compressible Navier–Stokes–Poisson system in the critical framework. J. Differential Equations (2019) 266: 6426-6458.
    [22] Global existence for the non-isentropic compressible Navier-Stokes-Poisson system in three and higher dimensions. Nonlinear Anal. Real World Appl. (2012) 13: 650-664.
    [23] Decay of the non-isentropic Navier–Stokes–Poisson equations. J. Math. Anal. Appl. (2013) 400: 293-303.
    [24] W. Wang and X. Xu, The decay rate of solution for the bipolar Navier–Stokes–Poisson system, J. Math. Phys., 55 (2014), 091502, 22 pp. doi: 10.1063/1.4894766
    [25] Decay of the Navier-Stokes-Poisson equations. J. Differential Equations (2012) 253: 273-297.
    [26] Pointwise estimates for bipolar compressible Navier–Stokes–Poisson system in dimension three. Arch. Rational Mech. Anal. (2017) 226: 587-638.
    [27] Optimal decay rate of the non-isentropic compressible Navier-Stokes-Poisson system in . J. Differential Equations (2011) 250: 866-891.
    [28] On the existence of solutions to the Navier-Stokes-Poisson equations of a two-demensional compressible flow. Math. Methods Appl. Sci. (2007) 30: 305-329.
    [29] F. Zhou and Y. Li, Convergence rate of solutions toward stationary solutions to the bipolar Navier–Stokes–Poisson equations in a half line, Bound. Value Probl., 2013 (2013), 22 pp. doi: 10.1186/1687-2770-2013-124
    [30] Large time behaviors of the isentropic bipolar compressible Navier–Stokes–Poisson system. Acta Math. Sci. Ser. B (Engl. Ed.) (2011) 31: 1725-1740.
  • Reader Comments
  • © 2021 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(2342) PDF downloads(214) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog