• 検索結果がありません。

In particu- lar, we obtain the optimal decay rates of the highest-order spatial derivatives of the velocity

N/A
N/A
Protected

Academic year: 2022

シェア "In particu- lar, we obtain the optimal decay rates of the highest-order spatial derivatives of the velocity"

Copied!
25
0
0

読み込み中.... (全文を見る)

全文

(1)

Electronic Journal of Differential Equations, Vol. 2020 (2020), No. 102, pp. 1–25.

ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu

OPTIMAL TIME DECAY RATES FOR THE COMPRESSIBLE NAVIER-STOKES SYSTEM WITH AND WITHOUT

YUKAWA-TYPE POTENTIAL

QING CHEN, GUOCHUN WU, YINGHUI ZHANG, LAN ZOU Communicated by Hongjie Dong

Abstract. We consider the time decay rates of smooth solutions to the Cauchy problem for the compressible Navier-Stokes system with and without a Yukawa- type potential. We prove the existence and uniqueness of global solutions by the standard energy method under small initial data assumptions. Further- more, if the initial data belong toL1(R3), we establish the optimal time decay rates of the solution as well as its higher-order spatial derivatives. In particu- lar, we obtain the optimal decay rates of the highest-order spatial derivatives of the velocity. Finally, we derive the lower bound time decay rates for the solution and its spacial derivatives.

1. Introduction

We consider the Cauchy problem of the compressible Navier-Stokes system with and without the Yukawa-type potential term in the whole spaceR3,

ρt+ div(ρu) = 0,

(ρu)t+ div(ρu⊗u) +∇P(ρ) +γρ∇ψ=µ∆u+ (µ+ν)∇divu,

−∆ψ+ψ=ρ−1,

(ρ, u)|t=0= (ρ0, u0)→(1,0) as |x| → ∞.

(1.1)

Here ρ = ρ(t, x), u = u(t, x), P = P(t, x) and ψ(t, x) represent the density, the velocity vector field of the fluid, the pressure and the potential force exerted in the fluid respectively, at timet≥0 and positionx∈R3. The Lam´e coefficientsµandν satisfyµ >0 and23µ+ν >0. And the constantγ∈Rmay be arbitrary and it is es- sential on its sign. Whenγ= 0, (1.1) reduces to compressible Navier-Stokes system, which describes the motion of a barotropic viscous compressible flow. Whenγ6= 0, (1.1) becomes the compressible Navier-Stokes equations with a Yukawa-potential, which is a simplified hydrodynamical model describing the nuclear matter [3, 8]. In this paper, we consider the compressible Navier-Stokes system with and without a Yukawa-potential. For technical consideration, we assume that the constantγ≥0 and the pressureP is some smooth function depending only onρandP0(ρ)>0.

2010Mathematics Subject Classification. 35Q30, 76N15.

Key words and phrases. Compressible flow; energy method; optimal decay rates.

c

2020 Texas State University.

Submitted February 2, 2020. Published September 29, 2020.

1

(2)

There are many important investigations on large time behavior of the solu- tions to the compressible Navier-Stokes system in multi-dimensional space, see [5, 10, 11, 12, 18, 20, 21, 23, 28, 27] and references therein. Matsumura and Nishida [20, 21] first proved the existence of the small global solutions inH3(R3) to compressible Navier-Stokes equations, and particularly, for the initial perturbation small inL1(R3)∩H3(R3), by employing time-decay properties of the linear system, the authors in [20] obtained the decay rate of the solution inL2-norm:

k(ρ−1, u)kL2(R3)≤C(1 +t)−3/4,

which is the same as for the heat equation with initial data inL1(R3). Later, for small initial disturbance in Hl(Rd)∩Wl,1(Rd) with l ≥ 4, Ponce [23] gave the optimal Lp (p ≥2) decay rates of the solutions and their first and second order derivatives. For the initial perturbation inH3(R3)∩Lp(R3) with 1≤p < 65, Duan, Liu, Ukai and Yang [6] proved the optimal convergence rates of the solution in Lq-norm with 2≤ q≤6 and its first order derivative in L2-norm. For the small data (ρ0−1, m0) with the momentum m=ρu in Hl(R3)∩B˙1,∞−s (R3) withl ≥4 and 0 ≤ s ≤ 1, Li and Zhang [17] studied the Cauchy problem for compressible Navier-Stokes system (1.1) withγ= 0 and obtained the following decay rate

k(ρ−1, m)kL2(R3)≤C(1 +t)−(34+2s).

Moreover, they established the lower bound of the time-decay rate for the global solution. For the small initial perturbation in k(ρ0−1, u0)kHl(R3) with l≥3 and the data bounded in ˙H−s(R3) with s ∈ [0,32), instead of resorting to the decay properties of the linear system, Guo and Wang [10] developed a general energy method and obtained the optimal decay rates of the higher-order spatial derivatives of solutions

k∇k(ρ−1, u)kL2(R3)≤C(1 +t)k+s2 , 1≤k≤l−1.

Recently, Danchin and Xu [5] studied theLp decay rates of the global solutions in the critical Lp framework and Xin and Xu [27] improved the result by removing some low frequencies conditions. Under the discontinuous initial data assumption, Hu and Wu [13] established the optimal convergence rates of the solutions with low regularity in Lp-norm with 2≤p≤ ∞and of the first order derivative of the velocity inL2-norm.

Compared with the compressible Navier-Stokes equations, there are few results on the system (1.1) withγ6= 0. For instance, Chikami [3] studied the existence and uniqueness of the solution in the critical space, and they also developed a blow-up criterion of the solution. In this paper, we first establish the global existence of the smooth solutions for the system (1.1), and then we will continue to address the optimal decay rates for the solutions. In particular, we can derive the optimal decay rate on the highest-order derivatives of the velocity.

The system (2.1) can be rewritten as

Ut=DU+N, U|t=0=U0,

with the solutionU = (ρ−1, u) and the matrix-valued differential operator Dhas the form

D=

0 −div

−P0(1)∇ −γ∇(1−∆)−1 µ∆ + (µ+ν)∇div

.

(3)

Thus the solution can be expressed as U(t) =E(t)U0+

Z t 0

E(t−τ)N(U(τ))dτ,

whereE(t) =etD is the solution semigroup. Due to the estimates on the semigroup and the energy estimates on the solution to the nonlinear problem (cf. [6, 17, 20]), it is difficult to show that the optimal decay rate of the high-order derivatives of the solution since the nonlinear term involves the derivatives. To improve the known results, motivated by [2, 4], we introduce Hodge decomposition to analyze the system (1.1). Hence the second equation of (1.1) can be divided into two systems. One is a mere heat equation on the “incompressible part”, whose decay rate is exponential; another one is a mixed system, which seems more complicated because of the nonlinear term. Fortunately, we find that the solution semigroup of the mixed system also keep good properties (the detail can be seen in the proof of Proposition 3.3). As a result, we can eventually derive the optimal decay rate of the highest order derivatives of the velocity.

Notation. In this paper,∇`with an integer`≥0 stands for the usual any spatial derivatives of order`. We useLp(R3) with 1≤p≤ ∞to denote the usualLpspaces with normk·kLp, andHs(R3) to denote the usual Sobolev spaces with normk·kHs. Furthermore, we use ˙Hs(R3) to denote the homogenous Sobolev spaces with norm k · kH˙s defined as

kfkH˙s =:kΛsfkL2 =k|ξ|sfˆkL2

with s∈ Rand here Λs is a Riesz potential operator of orders. We will employ the notationa.b to mean that a≤Cbfor a universal constant C >0 that only depends on the parameters coming from the problem. And Ci(i = 1,2,3,4) will also denote some positive constants depending only on the given problems.

Our main results are stated in the following theorems.

Theorem 1.1. Assume that k(ρ0−1, u0)kHl with an integerl ≥3 is sufficiently small. Then there exists a unique global solution (ρ(t, x), u(t, x), ψ(t, x)) to the initial value problem (1.1)such that

k(ρ−1, u)(t)k2Hl+kψ(t)k2Hl+2

+ Z t

0

(k∇ρ(τ)k2Hl−1+k∇u(τ)k2Hl+k∇ψ(τ)k2Hl+1)dτ .k(ρ0−1, u0)k2Hl.

(1.2)

If further k(ρ0−1, u0)kL1<+∞, then fork= 0,1, . . . , l−1,

k∇k(ρ−1)(t)kL2+k∇kψ(t)kH2 .(1 +t)−(34+k2), (1.3) and fork= 0,1, . . . , l,

k∇ku(t)kL2 .(1 +t)−(34+k2). (1.4) Remark 1.2. The results in Theorem 1.1 indicate that the optimal time decay rates are same for the solutions of the Navier-Stokes system and what with a Yukawa-type potential.

It is worth noting that the optimal time decay rates of the highest-order spa- tial derivatives of the velocity are obtained. By comparison, the optimal time decay rates of the compressible Navier-Stokes equations and Navier-Stokes-Poisson

(4)

equations, see [6, 7, 10, 15, 16, 26] and the references therein for instance, can be obtained but except the highest-order one. This is because of the decomposition on the system.

To obtain the optimal time decay rates of the higher-order derivatives of the solution, we can represent the spatial derivatives of the solutions to the equation Ut = BU+G with the initial dataU|t=0 = U0 which follows from the Duhamel principle as (cf. [25])

kU =∇kS(t)U0+ Z t/2

0

kS(t−τ)G(τ)dτ+ Z t

t/2

rS(t−τ)∇k−rG(τ)dτ, (1.5) whereS(t) :=etB is the solution semigroup and 0≤r≤k.

Note that the decay rates for the solutions and their derivatives above are op- timal. Indeed, we shall establish the lower bound of the time decay rates for the global solution.

Theorem 1.3. Besides the assumptions of Theorem 1.1, assume that the Fourier transform (F[ρ0−1],F[m0]) with m0 = ρ0u0 satisfies |F[ρ0−1](ξ)| > c0K0, and F[m0](ξ) = 0 for 0 ≤ |ξ| 1 with a positive constant c0 and K0 = k(ρ0− 1, u0)kL1∩Hl. Then, the global solution (ρ, u, ψ) given by Theorem 1.1 satisfies fort≥t0 witht0>0 a sufficiently large time such that fork= 0, . . . , l−1,

c1K0(1 +t)−(34+k2)≤min

k∇k(ρ−1)(t)kL2,k∇ku(t)kL2,k∇kψ(t)kH2

≤C(1 +t)−(34+k2),

(1.6) and

c1K0(1 +t)−(34+k2)≤ k∇lu(t)kL2 ≤C(1 +t)−(34+k2), (1.7) wherec1 is a positive constant independent of time.

Remark 1.4. Compared to the lower bound of the time-decay rate for the solu- tion obtained in [17], we can also get the lower-time-decay-rate for the high-order derivatives of the solution as well as the highest-order derivatives of the velocity.

The rest of this article is organized as follows. In Section 2, we reformulate the problem and state the equivalent theorem and propositions. In Section 3, we use the decomposition of the momentum to analyze the linearized system and establish the linearL2 decay estimates. In Section 4, we prove the global existence Proposition 2.2. In Section 5, we prove the optimal time decay rates Proposition 2.3 and the lower time decay rates Proposition 2.5 respectively.

2. Reformulated system

Denoting%=ρ−1, then we rewrite (1.1) in a perturbation form as

%t+ divu=N1,

ut+P0(1)∇%+γ∇ψ−µ∆u−(µ+ν)∇divu=N2,

−∆ψ+ψ=%,

(%, u)|t=0= (%0, u0) = (ρ0−1, u0),

(2.1)

where the nonlinear terms are

N1=−div(%u), (2.2)

(5)

N2=−u· ∇u− P0(ρ)

ρ −P0(1)

∇%−µ

ρ%∆u−µ+ν

ρ %∇divu. (2.3) For anyT >0, we define the solution space by

X(0, T) =n

(%, u, ψ) :%∈C0(0, T;Hl(R3))∩C1(0, T;Hl−1(R2)), u∈C0(0, T;Hl(R3))∩C1(0, T;Hl−2(R2)),

ψ∈C0(0, T;Hl+2(R3))∩C1(0, T;Hl+1(R2)),

∇%∈L2(0, T;Hl−1(R3)),∇u∈L2(0, T;Hl(R3)),

∇ψ∈L2(0, T;Hl+1(R3))o ,

(2.4)

and the solution norm by χ2(T) = sup

0≤t≤T

nk(%, u)(t)k2Hl+kψk2Hl+2

+ Z T

0

(k∇%(t)k2Hl−1+k∇u(t)k2Hl+k∇ψ(t)k2Hl+1)dto .

(2.5)

We now state a local existence theorem for the system (2.1), which can be es- tablished by a standard contraction mapping argument; we may refer to [15].

Proposition 2.1 (local existence). Let (%0, v0)∈Hl(R3)for an integerl≥3 and inf

x∈R3{%0+ 1}>0. (2.6)

Then there exists a positive constant T0 depending on χ(0) such that the problem (2.1)has a unique solution(%, u, ψ)∈X(0, T0)satisfying

inf

x∈R3,0≤t≤T0

{%(t, x) + 1}>0 and χ(T0)≤2χ(0). (2.7) It is easy to check that the global existence part of Theorem 1.1 is equivalent to the following proposition.

Proposition 2.2 (Global existence). Assume k(%0, u0)kl with an integer l≥3 is sufficiently small. Then there exists a unique global solution(ρ(t, x), u(t, x), ψ(t, x)) to the initial value problem (2.1)such that

k(%, u)(t)k2Hl+kψ(t)k2Hl+2

+ Z t

0

(k∇%(τ)k2Hl−1+k∇u(τ)k2Hl+k∇ψ(τ)k2Hl+1)dτ ≤Ck(%0, u0)k2l. (2.8) To obtain the lower time decay rate for the system (1.1), we consider the following linearized system, which is equivalent to (1.1):

%t+ divm= 0,

mt+P0(1)∇%+γ∇ψ−µ∆m−(µ+ν)∇divm=N,

−∆ψ+ψ=%,

(%, m)|t=0= (%0, m0) = (ρ0−1, ρ0u0),

(2.9)

(6)

where

N =: divF =: div

(−P(1 +%) +P(1) +P0(1)%)I3 +γ∇ψ⊗ ∇ψ−γ

2(|ψ|2+|∇ψ|2)I3−m⊗m 1 +%

−µ∇ %m 1 +%

−(µ+ν) div %m 1 +%

I3

.

(2.10)

For simplicity, we will also use the system (2.9) to analyze the upper decay rate for the solutions in the following proposition.

Proposition 2.3 (Optimal time-decay-rate). Under the assumptions in Proposi- tion 2.2, and thatk(%0, u0)kL1 <+∞, for k= 0, . . . , l−1, we have

k∇k%(t)kL2+k∇kψ(t)kH2.(1 +t)−(34+k2), (2.11) k∇l%(t)kL2+k∇l+2ψ(t)kL2.(1 +t)−(34+l−12 ), (2.12) and fork= 0, . . . , l, we have

k∇km(t)kL2.(1 +t)−(34+k2). (2.13) Remark 2.4. Regarding Proposition 2.3 we have the following observations.

• k(%0, u0)kL1 <+∞andkρ0kL <+∞implykm0kL1 <+∞(cf. [9]).

• By (2.11) and (2.13), we can check that under the assumptions of Proposi- tion 2.3, (1.4) holds fork= 0, . . . , l.

Proposition 2.5(lower-time decay rate). Assume that the conditions in Proposi- tions 2.2 and 2.3 hold, |F[%0](ξ)|> c0K0, andF[m0](ξ) = 0for0≤ |ξ| 1. Then fort ≥t0 with t0 >0 a sufficiently large time and fork = 0, . . . , l−1, the global solution (ρ, m, ψ)given by Proposition 2.2 satisfies

c2K0(1 +t)−(34+k2)≤min

k∇k(%, m)(t)kL2,kψ(t)kH2 ≤C(1 +t)−(34+k2), (2.14) c2K0(1 +t)(34+l2)≤ k∇lm(t)kL2 ≤C(1 +t)(34+2l), (2.15) wherec2 is a positive constant independent of time.

By using the estimates on the upper decay rates of the solution, we can conclude from (2.14)–(2.15) that (1.7) holds fork= 0, . . . , l.

3. Spectral analysis and linear L2 estimates

In this section, we focus on the decay rate of the solution to the linear system

%t+ divm= 0, mt+ P0(1)∇+γ∇(1−∆)−1

%−µ∆m−(µ+ν)∇divm= 0, (%, m)|t=0= (%0, m0) = (ρ0−1, ρ0u0).

(3.1)

Motivated by [2], we decompose the momentum m to analyze the above system (3.1) similarly as Hodge decomposition of the vector field, the system (3.1) can be transformed into two systems. One is a mere heat equation on the “incompressible part”, and another one has distinct eigenvalues. Let n = Λ−1divm and M =

(7)

Λ−1curlm (with curlz = (∂x2z3−∂x3z2, ∂x3z1−∂x1z3, ∂x1z2−∂x2z1)t), then we can rewrite (2.9) as follows

%t+ Λn= 0,

nt−Λ(P0(1) +γ(1−∆)−1)%−(2µ+ν)∆n= 0, Mt−µ∆M = 0,

(%, n, M)|t=0= (%0, n0, M0) = (%0−1divm0−1curlm0).

(3.2)

Indeed, as the definition ofnandM, and relation

m=−Λ−1∇n−Λ−1curlM (3.3)

involve pseudo-differential operators of degree zero, the estimates in the space Hl(R3) for the original functionm can be derived fromnandM.

Now, we study the time decay rates of the solutions for the system (3.1). It follows immediately that the convergence rate of M is exponential in any norm.

Hence, it suffices to consider the system

%t=−Λn, nt= Λ P0(1) +γ(1−∆)−1

%+ (2µ+ν)∆n, (%, n)|t=0= (%0, n0).

(3.4) In terms of the semigroup theory, by denotingV = (%, n)t, we may express (3.4) as,

Vt=BV,

V|t=0=V0 (3.5)

with

B=

0 −Λ

Λ P0(1) +γ(1−∆)−1

(2µ+ν)∆

Applying the Fourier transform to the system (3.5), we have Vbt=A(ξ)V ,b

Vb|t=0=Vb0,

(3.6) whereVb(t, ξ) =FV(t, x),ξ= (ξ1, ξ2, ξ3)tandA(ξ) is defined by

A(ξ) =

0 −|ξ|

P0(1) + 1+|ξ|γ 2

|ξ| −(2µ+ν)|ξ|2

!

. (3.7)

The eigenvalues of the matrixA(ξ) are computed from the determinant det(A(ξ)−λI) =λ2+ (2µ+ν)|ξ|2λ+

P0(1) + γ 1 +|ξ|2

|ξ|2= 0, (3.8) which implies the eigenvalues of the matrixAcan be expressed as

λ±(|ξ|) =− µ+ν 2

|ξ|2± r

µ+ν 2

2

|ξ|4− P0(1) + γ 1 +|ξ|2

|ξ|2. (3.9) The semigroupS(t) =etAcan be decomposed into

etA(ξ)=eλ+tP+(ξ) +eλtP(ξ), (3.10) where the projectorP±(ξ) is

P±(ξ) = A(ξ)−λI

λ±−λ . (3.11)

(8)

To estimate the semigroupetAinL2frame, we analyze the asymptotical expan- sions of λ±, P± and etA(ξ) for both lower and higher frequencies. From [21], we have the following lemma by careful computation.

Lemma 3.1. (a) For|ξ| 1, the spectral has the Taylor series expansion λ± =− µ+ν

2

|ξ|2+O(|ξ|4)±i (P0(1) +γ)1/2|ξ|+O(|ξ|3)

. (3.12)

(b) For|ξ| 1, the spectral has the Laurent expansion λ+=− P0(1)

2µ+ν +O(|ξ|−2), λ=−(2µ+ν)|ξ|2+ P0(1)

2µ+ν +O(|ξ|−2).

(3.13)

By (3.10)–(3.11) and Lemma 3.1, we obtain the following estimates for the so- lutionVb(t, ξ) to the system (3.6).

Lemma 3.2. (a) For|ξ| 1, we have

|%|,b |bn|.e−(µ+ν2)|ξ|2t(|%b0|+|nb0|), (3.14) (b) For|ξ| 1, we have

|%|b .e−Rt |%b0|+|ξ|−1|bn0|

, (3.15)

|bn|.|ξ|−1e−Rt|%b0|+

e−(µ+ν2)|ξ|2t+|ξ|−2e−Rt

|bn0| (3.16) for some positive constantR.

Proof. By the formula (3.10)–(3.11), we can calculate the semigroupS as follows.

S(t, ξ)

= (Sij(t, ξ))2×2 (3.17)

=

g1+, λ) −|ξ|g2+, λ)

|ξ| P0(1) +1+|ξ|γ 2

g2+, λ) g1+, λ)−(2µ+ν)|ξ|2g2+, λ)

, where

g1+, λ) =λ+eλt−λeλ+t

λ+−λ , g2+, λ) = eλ+t−eλt λ+−λ .

From the expansions (3.12) and (3.13) of λ±, we can estimategi+, λ) (i= 1,2) as follows

g1+, λ)

=

























e−(µ+ν2)|ξ|2

(µ+ν2)|ξ|2

(P0(1)+γ)1/2|ξ|+O(|ξ|3)sin

(P0(1) +γ)1/2|ξ|

+O(|ξ|3) t

+ cos

(P0(1) +γ)1/2|ξ|+O(|ξ|3) t

, if|ξ| 1,

O(1)e(−(2µ+ν)|ξ|2 +O(1))t+ (2µ+ν)|ξ|2+O(1)

e(−

P0(1)

2µ+ν+O(|ξ|−2 ))t

(2µ+ν)|ξ|2+O(1) ,

if|ξ| 1,

(3.18)

(9)

and

g2+, λ) =





sin

(P0(1)+γ)1/2|ξ|+O(|ξ|3)

t

(P0(1)+γ)1/2|ξ|+O(|ξ|3) e−(µ+ν2)|ξ|2, |ξ| 1,

e

P0(1)

2µ+ν+O(|ξ|−2 )

t−e −(2µ+ν)|ξ|2 +O(1)

t

(2µ+ν)|ξ|2+O(1) , |ξ| 1.

(3.19)

Hence we conclude that

|g1+, λ)|.

(e−(µ+ν2)|ξ|2t, |ξ| 1,

e−Rt, |ξ| 1 (3.20)

for a positive constantR, and

|g2+, λ)|.

(|ξ|−1e−(µ+ν2)|ξ|2t, |ξ| 1,

|ξ|−2e−Rt, |ξ| 1.

(3.21) Moreover, by delicate calculations, we have the estimate

|g1+, λ)−(2µ+ν)|ξ|2g2+, λ)|

.

(e−(µ+ν2)|ξ|2t, |ξ| 1, e−(µ+ν2)|ξ|2t+|ξ|−2e−Rt, |ξ| 1.

(3.22)

Now we represent the solution of (3.6) as

Vb(t, ξ) =etAVb0. (3.23) Therefore, by plugging (3.20)–(3.22) into the expression (3.17) of S(t), we obtain

(3.14)–(3.16).

As in [21], we obtain the decay rates for the solution (%, n, M) of the linear system (3.2) as follows.

Proposition 3.3. Assume that(%0, m0)∈Hl∩L1. Letn= Λ−1divmand M = Λ−1curlm. Then the solution (%, n, M)of the linear system (3.2)satisfies

(a)

k%k2L2 .(1 +t)−3/2k(%0, n0)k2L1+e−2Rtk(%0, n0)k2L2, (3.24) and for1≤k≤l,

k∇k%k2L2.(1 +t)−(32+k)k(%0, n0)k2L1

+e−2Rt

k∇k%0k2L2+k∇k−1n0k2L2

.

(3.25) (b) Fork= 0,1,

k∇knk2L2.(1 +t)(32+k)k(%0, n0)k2L1+e−2Rtk(%0, n0)k2L2, (3.26) and for2≤k≤l,

k∇knk2L2 .(1+t)−(32+k)k(%0, n0)k2L1+e−2Rt

k∇k−1%0k2L2+k∇k−2n0k2L2

, (3.27) and for0≤k≤l,

k∇kMk2L2 .(1 +t)−(32+k)kM0k2L1. (3.28)

(10)

Proof. We only prove (3.27). By Lemma 3.2, Plancherel theorem and Hausdorff- Young’s inequality, from (3.14) and (3.16) we have that for each 2≤k≤l and for someη >0,

k∇knk2L2=k|ξ|knkb 2L2

. Z

|ξ|<η

|ξ|2ke−(2µ+ν)|ξ|2t(|%b0|2+|bn0|2)dξ+ Z

|ξ|≥η

|ξ|2ke−(2µ+ν)|ξ|2t|bn0|2dξ +

Z

|ξ|≥η

e−2Rt(|ξ|k−1|%b0|+|ξ|k−2|bn0|)2dξ . k%b0k2L+knb0k2L

Z

|ξ|<∞

|ξ|2ke−(2µ+ν)|ξ|2tdξ +e−2Rt

Z

|ξ|≥η

(|ξ|k−1|%b0|+|ξ|k−2|bn0|)2

.(1 +t)−(32+k)k(%0, n0)k2L1+e−2Rt k∇k−1%0k2L2+k∇k−2n0k2L2

.

(3.29)

Finally, to obtain the lower decay rates for the solution, we also need the following decay rates for the solution (%, n, M) of the linear system (3.2).

Proposition 3.4. Assume that %0 ∈ Hl∩H˙−s and m0 ∈ L2∩H˙−s. Let n = Λ−1divm and M = Λ−1curlm. Then the solution(%, n, M) of the linear system (3.2)satisfies

(a)

k%k2L2 .(1 +t)−sk(%0, n0)k2H˙−s+e−2Rtk(%0, n0)k2L2, (3.30) k∇%k2L2.(1 +t)−(1+s)k(%0, n0)k2H˙−s+e−2Rt k∇%0k2L2+kn0kL2

. (3.31) (b) Fork= 0,1,

k∇knk2L2 .(1 +t)−(k+s)k(%0, n0)k2H˙−s+e−2Rtk(%0, n0)k2L2, (3.32) k∇kMk2L2.(1 +t)−(k+s)kM0k2H˙−s. (3.33) See [1] for a proof of this proposition, or use an argument similar to the one in Proposition 3.3.

4. Energy estimates

To prove Proposition 2.2, by the standard continuity argument, it suffices to derive the following a priori energy estimates.

Proposition 4.1 (a priori estimate). Let (%0, u0)∈Hl(R3)with an integerl≥3.

Suppose that (2.1)has a solution(%, u, ψ)∈X(0, T), whereT is a positive constant.

Then there exists a small constantδ >0, independent ofT, such that if sup

0≤t≤T

k(%, u)(t)kHl+kψ(t)k2Hl+2 ≤δ, (4.1) then for any t∈[0, T],

k(%, u)(t)k2Hl+kψ(t)k2Hl+2+ Z t

0

(k∇%(τ)k2Hl−1+k∇u(τ)k2Hl+k∇ψ(τ)k2Hl+1)dτ .k(%0, u0)k2Hl.

(4.2)

(11)

First from (2.1)3, we can easily deduce the following lemma and we omit the proof.

Lemma 4.2. Under the assumption of Proposition 4.1, for k= 0, . . . , l, it holds k∇kψkH2 ≈ k∇k%kL2. (4.3) Next we derive an energy estimates for (%, u).

Lemma 4.3. Under the assumption of Proposition 4.1, there exists a positive con- stantC1, such that

d dt

P0(1)

2 k%(t)k2L2+1

2kuk2L2

2kψk2H1+ µ 4C1

h∇%, ui +µ

2k∇u(t)k2L2+µP0(1)

16C1 k∇%(t)k2L2+ µγ

4C1k∇ψk2H1

≤Ck∇lu(t)k2L2.

(4.4)

Proof. From (4.1) and the Sobolev inequality, we obtain upper and lower bounds 1

2 ≤ρ≤ 3

2. (4.5)

Multiplying (2.1)1and(2.1)2withP0(1)%andurespectively, and then integrating the resulting equalities overR3, one has

d dt

P0(1)

2 k%(t)k2L2+1

2kuk2L2 +hγ∇ψ, ui+µk∇uk2L2+ (µ+ν)kdivuk2L2

=hN1, P0(1)%i+hN2, ui.

(4.6)

The terms in (4.6) can be estimated as follows. By using (2.1)1, (2.1)3, and inte- gration by parts, we have

hγ∇ψ, ui=hγψ,−divui

=hγψ, %t+ div(%u)i

=hγψ,(−∆ψ+ψ)ti+hγψ, %divu+∇%·ui

≥ γ 2

d

dtkψk2H1−CkψkL3(k%kL6kdivukL2+k∇%kL2kukL6)

≥ γ 2

d

dtkψk2H1−Cδ k∇%k2L2+k∇uk2L2

,

(4.7)

where the a priori assumption (4.1), H¨older’s inequality, Cauchy’s inequality and the Sobolev embedding theorem are used. In addition, by (4.5), we have

hN1, %i=h−div(%u), %i ≤Cδ k∇%k2L2+k∇uk2L2

, (4.8)

(12)

hN2, ui

=h−u· ∇u, ui+

− P0(ρ)

ρ −P0(1)

∇%, u +

µ∇ % ρ

· ∇u, u +µ

ρ%∇u,∇u +

(µ+ν)∇ % ρ

divu, u

+µ+ν

ρ %divu,divu

≤C

kukL3k∇ukL2kukL6+k%kL3k∇%kL2kukL6

+kukLk∇%kL2k∇ukL2+k%kLk∇uk2L2

≤Cδ(k∇%k2L2+k∇uk2L2).

(4.9)

Plugging (4.7)–(4.9) into (4.6) and using Cauchy’s inequality and the smallness of δ, we can deduce that

d dt

nP0(1)

2 k%(t)k2L2+1

2kuk2L2+γ 2kψk2H1

o +3µ

4 k∇uk2L2+ (µ+ν)kdivuk2L2

≤Cδk∇%k2L2.

(4.10) Next we shall deal with theL2-norm of∇%. By takingh∇(2.1)1, ui+h(2.1)2,∇%i, we have

d

dth∇%, ui+P0(1)k∇%(t)k2L2+hγ∇ψ,∇%i

=h∇(−divu+N1), ui+hµ∆u+ (µ+ν)∇divu+N2,∇%i.

(4.11) By using (2.1)3, we obtain the following estimates:

hγ∇ψ,∇%i=hγ∇ψ,∇(−∆ψ+ψ)i=γk∇ψk2H1, (4.12) h∇(−divu+N1), ui=hdivu+ div(%u),divui

≤ kdivuk2L2+k(%, u)kLk∇(%, u)kL2k∇ukL2

≤Ck∇uk2L2+Cδk∇%k2L2,

(4.13)

and

hµ∆u+ (µ+ν)∇divu+N2,∇%i

≤ P0(1)

4 k∇ρk2L2+Ck∇2uk2L2+Ck(%, u)kLk ∇%,∇u,∇2u

kL2k∇%kL2

≤ P0(1)

4 k∇ρk2L2+C k∇uk2L2+k∇luk2L2

+Cδ k∇%k2L2+k∇uk2L2

.

(4.14)

Thus plugging (4.12)–(4.14) into (4.11) yields d

dth∇%, ui+P0(1)

2 k∇%(t)k2L2+γk∇ψk2H1 ≤C1k∇uk2L2+Ck∇luk2L2, (4.15) where C1 is some positive number. Then the estimate (4.4) follows by taking the addition of (4.10) andµ/(4C1) times (4.15) and using the smallness ofδ.

(13)

Lemma 4.4. Under the assumption of Proposition 4.1, there exist two positive constants C2 andC3 such that

d dt

nP0(1)

2 k∇l−1%k2H1+1

2k∇l−1uk2H1

2k ∇l−1ψ,∇lψ k2H1

+ µ

4C2h∇l%,∇l−1uio +µ

2k∇lu(t)k2H1+µP0(1)

16C2 k∇l%(t)k2L2

+ µγ

4C2k∇lψk2H1≤0.

(4.16)

Proof. By taking derivatives withk=l−1 orl, we obtain 1

2 d dt

n

P0(1)k∇k%k2L2+k∇kuk2L2

o

+µk∇k+1u(t)k2L2+ (µ+ν)k∇kdivuk2L2

+hγ∇k∇ψ,∇kui

=hP0(1)∇kN1,∇k%i+h∇kN2,∇kui.

(4.17)

By using (2.1)1, (2.1)3, and Lemmas 4.2, 6.1–Lemma 6.4, we can estimate the terms in (4.17) as follows.

hγ∇k∇ψ,∇kui=hγ∇kψ,∇k(−divu)i

=hγ∇kψ,∇k(%t+ div(%u))i

=hγ∇kψ,∇k(−∆ψ+ψ)ti+hγ∇kψ,∇kdiv(%u)i

=γ 2

d

dtk∇kψk2H1− hγ∇k+1ψ,∇k(%u)i.

(4.18)

By

hγ∇lψ,∇l−1(%u)i

.k∇lψkL2 k∇l−1%kL6kukL3+k%kL3k∇l−1ukL6 .k(%, u)kH1k∇lψkL2k∇l(%, u)kL2

.δk∇l(%, u)k2L2,

(4.19)

and

|hγ∇l+1ψ,∇l(%u)i|.k∇l+1ψkL2 k%kL3k∇lukL6+k∇l%kL2kukL

k∇l%k2L2+k∇l+1uk2L2

, (4.20)

we can conclude that fork=l−1, l, hγ∇k∇ψ,∇kui ≥ γ

2 d

dtk∇kψk2H1−Cδ k∇l%k2L2+k∇k+1uk2L2

. (4.21)

Similarly we have

h∇kN1,∇k%i=h∇k(−∇%·u−%divu),∇k%i

= Z

R3

(divu)|∇k%|2

2 dx− h[∇k, u]· ∇ρ,∇kρi +k∇k(%divu)kL2k∇k%kL2,

(4.22)

(14)

where the commutator [∇k, f]g is defined in (6.6). This together with Z

R3

(divu)|∇l−1%|2

2 dx− h[∇l−1, u]· ∇ρ,∇l−1ρi +k∇l−1(%divu)kL2k∇l−1%kL2

. k∇(%, u)kL3/2k∇l−1(%, u)kL6+k%kL3k∇lukL2

k∇l−1%kL6

.δk∇l(%, u)k2L2,

(4.23)

and Z

R3

(divu)|∇l%|2

2 dx− h[∇l, u]· ∇ρ,∇lρi+k∇l(%divu)kL2k∇l%kL2

. k∇ukLk∇l%kL2+k∇%kL3k∇lukL6+k%kLk∇l+1ukL2

k∇l%kL2

k∇l%k2L2+k∇l+1uk2L2

(4.24)

implies that fork=l−1, l,

h∇kN1,∇k%i.δ

k∇l%k2L2+k∇k+1uk2L2

. (4.25)

Furthermore, we have from (2.3) that fork=l−1, h∇l−1N2,∇l−1ui

=h∇l−1 −u· ∇u

,∇l−1ui+

l−1

− P0(ρ)

ρ −P0(1)

∇%

,∇l−1u +

l−1 µ ρ%∇u

,∇lu +

l−1

∇ µ ρ%

· ∇u

,∇l−1u +

l−1 µ+ν

ρ %divu

,∇l−1divu +

l−1

∇ µ+ν ρ %

divu

,∇l−1u .

k(%, u)kL3k∇l(%, u)kL2+k∇l−1 P0(ρ) ρ

−P0(1), u

kL6k∇(%, u)kL3/2

k∇l−1ukL6

+

k%kLk∇lukL2+k∇l−1 % ρ

kL6k∇ukL3

k∇lukL2

+

k∇%kL3k∇lukL2+k∇l % ρ

kL2k∇ukL3

k∇l−1ukL6

k∇l%k2L2+k∇luk2L2

k∇l P0(ρ)

ρ −P0(1) kL2

+k∇l %

ρ

kL2

k∇lukL2

k∇l%k2L2+k∇luk2L2

,

(4.26)

(15)

where (6.9) is used, and fork=l, h∇lN2,∇lui

=h∇l(−u· ∇u),∇lui+

l−1 P0(ρ)

ρ −P0(1)

∇%

,∇ldivu +

l−1 µ ρ%∆u

,∇ldivu +

l−1 µ+ν

ρ %∇divu

,∇ldivu .

kukL3k∇l+1ukL2+k∇lukL6k∇ukL3/2

k∇lukL6

+

k%kLk∇l%kL2+k∇l−1

P0(ρ)

ρ −P0(1)

kL6k∇%kL3

+k%kLk∇l+1ukL2+k∇l−1 % ρ

kL6k∇2ukL3

k∇l+1ukL2

.k(%, u)kH3k ∇l%,∇l+1u

kL2k∇l+1ukL2

.δ k∇l%k2L2+k∇l+1uk2L2

.

(4.27)

Therefore, from (4.26) and (4.27) we conclude that fork=l−1, l, h∇kN2,∇kui.δ k∇l%k2L2+k∇k+1uk2L2

. (4.28)

Hence plugging (4.21), (4.25) and (4.28) into (4.17) yields that fork=l−1, l, 1

2 d dt

P0(1)k∇k%k2L2+k∇kuk2L2+γk∇kψk2H1 +3µ

4 k∇k+1u(t)k2L2

≤Cδk∇l%k2L2.

(4.29) Here the smallness ofδis used again.

Now we turn to estimate theL2-norm of∇l%. By taking the derivative, we have d

dth∇l%,∇l−1ui+P0(1)k∇l%(t)k2L2+hγ∇lψ,∇l%i

=h∇l(−divu+N1),∇l−1ui+h∇l−1(µ∆u+ (µ+ν)∇divu+N2),∇l%i.

(4.30) By using (2.1)3 and Lemmas 6.1–6.4, we obtain the following estimates.

hγ∇lψ,∇l%i=hγ∇lψ,∇l(−∆ψ+ψ)i=γk∇lψk2H1, (4.31) h∇l(−divu+N1),∇l−1ui

=h∇l−1(divu+ div(%u)),∇l−1divui

≤ k∇l−1divuk2L2+k(%, u)kLk∇l(%, u)kL2k∇lukL2

≤Ck∇luk2L2+Cδk∇l%k2L2,

(4.32)

and as in the proof of (4.26) and (4.27),

h∇l−1(µ∆u+ (µ+ν)∇divu+N2),∇l%i

≤P0(1)

4 k∇l%k2L2+Ck∇l+1uk2L2

+Ck(%, u)kH3k ∇l%,∇lu,∇l+1u

kL2k∇l%kL2

≤3P0(1)

8 k∇l%k2L2+Ck∇luk2H1.

(4.33)

参照

関連したドキュメント

The torsion free generalized connection is determined and its coefficients are obtained under condition that the metric structure is parallel or recurrent.. The Einstein-Yang

It is suggested by our method that most of the quadratic algebras for all St¨ ackel equivalence classes of 3D second order quantum superintegrable systems on conformally flat

[11] Karsai J., On the asymptotic behaviour of solution of second order linear differential equations with small damping, Acta Math. 61

Keywords: continuous time random walk, Brownian motion, collision time, skew Young tableaux, tandem queue.. AMS 2000 Subject Classification: Primary:

This paper develops a recursion formula for the conditional moments of the area under the absolute value of Brownian bridge given the local time at 0.. The method of power series

Abstract The representation theory (idempotents, quivers, Cartan invariants, and Loewy series) of the higher-order unital peak algebras is investigated.. On the way, we obtain

Then it follows immediately from a suitable version of “Hensel’s Lemma” [cf., e.g., the argument of [4], Lemma 2.1] that S may be obtained, as the notation suggests, as the m A

Abstract The classical abelian invariants of a knot are the Alexander module, which is the first homology group of the the unique infinite cyclic covering space of S 3 − K ,