ebook img

Global Wellposeness for the 3D inhomogeneous incompressible Navier-Stokes equations PDF

0.21 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Global Wellposeness for the 3D inhomogeneous incompressible Navier-Stokes equations

GLOBAL WELLPOSEDNESS FOR THE 3D INHOMOGENEOUS INCOMPRESSIBLE NAVIER-STOKES EQUATIONS WALTER CRAIG, XIANGDI HUANG, AND YUN WANG 3 1 0 2 n Abstract. This paper addresses the three-dimensional Navier-Stokes equations for an in- a compressiblefluidwhosedensityispermittedtobeinhomogeneous. Weestablishatheorem J of global existence and uniqueness of strong solutions for initial data with small H˙ 21-norm, 0 which also satisfies a natural compatibility condition. A key point of the theorem is that 3 the initial density need not be strictly positive. ] Keywords: inhomogeneous incompressible fluids, strong solutions, vacuum. P AMS: 35Q35,35B65, 76N10 A . h 1. Introduction t a m In a number of important applications to fluid mechanics, such as in geophysical fluid dynamics, the Navier-Stokes equations are called upon to describe situations in which a [ fluid is inhomogeneous with respect to density, however it is essentially incompressible to a 1 great extent. Such cases occur within the outer core and mantle of the earth, for example, or v 5 else within the atmosphere in a regime of flows for which dynamic effects of compressibility 5 do not play a principle rˆole. Ignoring any possible buoyancy or other body forces, flows of 1 7 this character satisfy the following system of inhomogeneous incompressible Navier-Stokes 1. equations; 0 ∂ ρ+div(ρu) = 0 , 3 t 1 (1.1)  ∂ (ρu)+div(ρu u) µ∆u+ P = 0 , t :  ⊗ − ∇ v div u = 0 . i X It is a standard mathematical model to pose the initial and boundary value problem, giving r the following data; a (ρ,u) = (ρ ,u ) in Ω , (1.2) |t=0 0 0 u = 0 on ∂Ω , u(x, ) 0 as x . · → | | → ∞ Here ρ, u and P denote the density, velocity and pressure of the fluid respectively, and µ is the positive viscosity, which is assumed to be a constant. In this paper, we consider Ω is either a smooth bounded domain in R3 or the whole space R3. For the initial density with positive lower bound, the inhomogeneous equations (1.1) have been studied in the sixties and seventies by the Russian school, see [4,21,24] and the many references that they contain. It was proved that a unique strong solution exists locally for The research of WC is supported in part by a Killam Research Fellowship, the Canada Research Chairs Program and NSERC through grant number 238452–11. The research of YW is supported in part by a Canada Research Chairs PostdoctoralFellowship at McMaster University. 1 2 WALTER CRAIG,XIANGDIHUANG,AND YUNWANG arbitrary initial data. Moreover, these papers also establish global wellposedness results for small solutions in dimension N 3, while for the two dimensional case they establish ≥ the existence of large strong solutions. More recently, there have been many subsequent contributions to this theory, obtaining global wellposedness for initial data belonging to certain scale invariant spaces, see for example [1,13,18]. On the other hand, for initial data which permits regions of vacuum, i.e. regions where the density ρ vanishes on some set, the problem becomes much more involved. Authors including DiPerna and Lions [15,25] prove the global existence of weak solutions to (1.1) in any space dimension, see also [14,27] and the references in their work. As pointed in [12], the major difficulty for existence of strong solutions which admit regions of vacuum is the lack of an appropriate estimate for ∂ u, since ∂ u only appears in the momentum equation t t with a possibly degenerate coefficient ρ. One way to avoid this is to estimate ∂ u instead, t ∇ for which one pays the price of being required to impose a compatibility condition. In fact, Choe-Kim [12] prove the following theorem, Theorem 1.1. Assume that the initial data (ρ ,u ) satisfies 0 0 (1.3) 0 ρ L23 H2 , u D1,2 D2,2 , ≤ 0 ∈ ∩ 0 ∈ 0 ∩ and the compatibility condition (1.4) µ∆u P = √ρ g , div u = 0 in Ω , 0 0 0 0 0 −∇ for some (P ,g ) D1,2 L2. Then there exists a time T > 0 and a unique strong solution 0 0 ∈ × (ρ,u,P) to the initial boundary value problem (1.1)-(1.2) satisfying ρ L∞(0,T;L23 H2), u C([0,T] ;D1,2 D2,2) L2(0,T;D3,2) , ∈ ∩ ∈ 0,σ ∩ ∩ ∂ u L2(0,T;D1,2) , and √ρ∂ u L∞(0,T;L2) . t ∈ 0,σ t ∈ It is also interesting to study the regularity criterion for such strong solutions, see [11,22] and its references. In particular, Kim [22] proved that if T∗ is the blowup time of a local strong solution, then necessarily T∗ 2 3 (1.5) u(s) p ds = , for any (p,q) , + = 1 , 3 < q , Z k kLq,∞ ∞ p q ≤ ∞ 0 where Lq,∞ is the usual weak Lebesgue space. This is effectively a criterion of regularity, and is similar to the Serrin criterion [26] for a priori regularity condition for the homogeneous Navier-Stokes equations. However, like the usual homogeneous Navier-Stokes equations, the question of global exis- tence of strong solutions remains open for large initial data. A partial result due to Kim [22] asserts that if u0 D1,2 is small enough, then a strong solution exists globally in time and is k k unique. In this paper, we generalize this result to the case of data within a class of critical Sobolev spaces. There is a scaling invariance of the system (1.1), namely if (ρ,u,P) is a solution associated with the initial data (ρ ,u ), then 0 0 (ρ , u ,P ) = (ρ(λ2t,λx) ,λu(λ2t,λx),λ2P(λ2t,λx)) λ λ λ INHOMOGENEOUS INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 3 is a solution to (1.1)-(1.2) associated with the initial data (ρ (λx),λu (λx)). A scale invari- 0 0 ant function space H∗ is one for which (ρλ,uλ) H∗ = (ρ1,u1) H∗. For example, scaling k k k k invariance implies that kukH˙ 12 = kuλkH˙ 12 , which is our principle example. Our main result in this paper is that if ku0kH˙ 21 is small enough, then a strong solution exists globally in time. More precisely, Theorem 1.2. Assume the conditions in Theorem 1.1. Then there exists a small constant ε depending on ρ¯= ρ0 L∞, µ and the domain Ω, such that if k k (1.6) ku0kH˙ 21 ≤ ε , then the unique local strong solution constructed in Theorem 1.1 exists globally in time. Moreover, this global strong solution satisfies the following decay properties: −1 (1.7) u L2 Mt 2 , k∇ k ≤ where M depends on µ, Ω, ρ¯, and ρ u 2dx. 0 0 | | R The global result above relies on the smallness of the scale invariant norm of the initial data. In terms of the well known result of Fujita & Kato [16] for the homogeneous Navier- Stokes equations, our result can be considered as its generalization to the inhomogeneous case. The homogeneous Navier-Stokes equations have been a much more active topic of inquiry in this direction, see [9,10,23] and their references. The key ideas leading to these results come from a variety of refined estimates of Stokes semigroup combined with the contraction principle for a Picard iteration scheme. However, in the present work one can not directly apply the same methods, due to the presence of the additional factor ρ before the term ∂ u. Instead we pursue a different strategy of proof, based on the method of energy t estimates, and using the parabolic property of the system of equations. The main idea goes as follows: we first assume that the scale-invariant quantity k∇ukL4tL2x is less than an a priori constant bound2; then using thisassumption we thenprove, under thesmallness assumption on the initial data, that in fact the quantity is less than 1. Since k∇ukL4tL2x is initially 0, which is less than 1, then it remains less than 1. On the other hand, the boundedness of k∇ukL4tL2x implies the local solution can be extended, according to some a priori estimates. The proof in this paper is inspired by a similar result for 3D compressible Navier-Stokes equations due to Huang, Li & Xin [19], in which they proved the global wellposedness of 1 classical solutions with small initial H2+ε-norm. Remarks: Compared to the global existence results in [1,13,18], for which the initial density isrequired tobeasmallperturbationofaconstant, ourresult doesnotrequireanassumption of smallness on the density variations, and in fact it allows for the presence of regions of vacuum. In the case of periodic boundary conditions on a torus Td, the same result can be proved if the intial data has zero momentum, following the strategy as for the case of a bounded domain. Indeed, thesameconclusionholdsforclassicalsolutionsiftheinitialdataareregular enough and satisfy some higher-order compatibility conditions. The proof of this statement is almost the same except that some additional high-order estimates are required. These 4 WALTER CRAIG,XIANGDIHUANG,AND YUNWANG estimates have been given in [22]. In the case in which Ω is a smooth bounded domain in R2, it is already known that global strong solutions exist without a smallness assumption; this was proved first for the case without vacuum [4] and subsequently for the case with vacuum [20]. Our result holds in particular in the case of a constant density ρ = 1. It therefore gives a new proof of the well known Fujita-Kato theorem [16] , which uses only energy estimate methods in the argument. Theorem 1.3. Assume u H1(R3) and div u = 0. There exists a positive constant ε, such 0 0 ∈ that if ku0kH˙ 21 ≤ ε, then there is a global strong solution to the homogeneous Navier-Stokes equations with the property that u C([0, );H1) L2 (0, ;H2) . ∈ ∞ ∩ loc ∞ This article is organized as follows: section 2 sets the notation, and contains several definitions and basic lemmas. In section 3 we give the proof for the case Ω = R3, while the proof for the case of a bounded domain is presented in section 4. 2. Notation and interpolation lemmas The homogeneous and inhomogeneous Sobolev spaces are defined in the standard way. For 1 r and k N, ≤ ≤ ∞ ∈ Lr = Lr(Ω), Dk,r = u L1 : ku < , u = ku , { ∈ loc k∇ kLr ∞} k kDk,r k∇ kLr Wk,r = Lr Dk,r , Hk = Wk,2 , Dk,r = C∞ closure in the norm of Dk,r . ∩ 0 0 In particular when Ω is a bounded domain then D1,2 = H1. Set 0 0 ∞ ∞ C = u C : div u = 0 , 0,σ { ∈ 0 } and set D1,2 = C∞, with the closure taken in the norm of D1,2. The fractional-order σ 0,σ homogeneous Sobolev space H˙ s(R3) are defined as the space of tempered distributions u over R3 for which the Fourier transform u belongs to L1 (R3) and which satisfy F loc u 2 := ξ 2s u(ξ) 2dξ < . k kH˙s(R3) Z | | |F | ∞ R3 And H˙ s(Ω) is the restriction of H˙ s(R3) to the domain Ω. We refer to H˙ s(Ω) by H˙ s, inde- pendent of whether Ω is the whole space or not. SomewellknowninterpolationresultswhichrelateLorentzspacesandtheclassicalSobolev spaces are presented, which can be found in [8,28,29]. For every 0 < θ < 1, 1 q , and ≤ ≤ ∞ normed spaces X ,X , we use the notation [X ,X ] to denote the real interpolation space 0 1 0 1 θ,q between X and X . 0 1 Lemma 2.1. Let s,s ,s R. If s = (1 θ)s +θs with some 0 < θ < 1, then 0 1 0 1 ∈ − [H˙ s0,H˙ s1] = H˙ s. θ,2 INHOMOGENEOUS INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 5 And the following inclusion relation holds, if 1 q q , then 1 2 ≤ ≤ ≤ ∞ [H˙ s0,H˙ s1] [H˙ s0,H˙ s1] . θ,q1 ⊆ θ,q2 Lemma 2.2. Let 0 < θ < 1, 1 < p ,p ,p < , 1 q ,q ,q , and 1 = 1−θ + θ , then 0 1 ∞ ≤ 0 1 ≤ ∞ p p0 p1 (2.8) [Lp0,q0,Lp1,q1] = Lp,q . θ,q This identification of the interpolation spaces (2.8) is also valid if Lp1,q1 is replaced by by L∞ (Lp0,q0, Lp1,q1 and Lp,q are Lorentz spaces, whose definition can be found in [8]). When q = p, Lp,p is the Sobolev Lp space. The above interpolation lemmas clearly generalize to vector-valued Lorentz spaces. For example, Lemma 2.3. Let 0 < θ < 1, 1 < p ,p < , 1 q ,q ,q , and 1 = 1−θ + θ . Then 0 1 ∞ ≤ 0 1 ≤ ∞ p p0 p1 (2.9) [Lp0,q0(0,T;L2),Lp1,q1(0,T;L2)] = Lp,q(0,T;L2) . θ,q The statement (2.9) is also valid after replacing Lp1,q1(0,T;L2) by L∞(0,T;L2). Bounds for linear operators on interpolation spaces are the concern of the next lemma. Lemma 2.4. Suppose that E ,E ,F ,F are Banach spaces and 0 < θ < 1, 1 q , 0 1 0 1 ≤ ≤ ∞ and that A is a linear operator from E + E into F + F . If A maps E into F , with 0 1 0 1 0 0 Aw L w for all w E , and maps E into F , with Aw L w for all k kF0 ≤ 0k kE0 ∈ 0 1 1 k kF1 ≤ 1k kE1 w E . Then A is a continous linear operator from [E ,E ] into [F ,F ] and one has 1 0 1 θ,q 0 1 θ,q ∈ Aw L1−θLθ w . k k[F0,F1]θ,q ≤ 0 1k k[E0,E1]θ,q In this paper, the constants C,C ,C ,C ,C ,C , may depend upon µ,ρ¯,Ω and as 1 2 3 4 5 ··· conventional in analysis, they may change from line to line. 3. Proof of Theorem 1.2 for Ω = R3 In this section, we focus on the case Ω = R3. 3.1. A Priori Estimates. This subsection establishes several a priori estimates for strong solutionstotheCauchyproblem, whichwillplayakeyroleinextending localstrongsolutions to global ones. Given a strong solution (ρ,u,P) on R3 [0,T], define × (3.10) A(t) = u L4(0,t; L2) , 0 t T . k∇ k ≤ ≤ Theorem 3.1. Under the assumptions of Theorem 1.1, there exists a positive constant ε depending only on µ and ρ¯, such that if ku0kH˙ 21 ≤ ε, and (ρ,u,P) is a strong solution to (1.1)-(1.2), satisfying (3.11) A(T) 2, ≤ then it in fact holds that A(T) 1. ≤ The remainder of this subsection consists in proving this key result. 6 WALTER CRAIG,XIANGDIHUANG,AND YUNWANG Lemma 3.2. Let (ρ,u,P) be a strong solution of (1.1)-(1.2). Then for all 0 t T, ≤ ≤ (3.12) ρ(t) L∞ = ρ0 L∞ = ρ¯ . k k k k Proof. The mass equation is in fact a transport equation, owing to the fact that div(u) = 0, (cid:3) from which (3.12) follows. Lemma 3.3. Let (ρ,u,P) be a strong solution of (1.1)-(1.2). Suppose that A(T) 2, ≤ then there exists some constant C depending on µ,ρ¯, such that 1 (3.13) k∇ukL4(0,T;L2) ≤ C1ku0kH˙ 12 , and it holds that (3.14) sup t u(t) 2 C ρ u 2dx. k∇ kL2 ≤ Z 0| 0| t∈[0,T] Note that C does not depend on T. 1 Proof. First, let’s consider the following linear Cauchy problem for (w,P˜), ˜ ρ∂ w µ∆w +(ρu )w+ P = 0 , t − ·∇ ∇  (3.15) div w = 0 ,   w(x,0) = w (x) . 0   Supposethatw satisfiestheconditionsassumedforu asprescribedinTheorem1.1, thenthe 0 0 existence anduniqueness ofstrongsolutionto(3.15)hasbeenproved in[12]. Straightforward energy estimates tell that 1 T 1 (3.16) ρ w(T) 2dx+µ w 2 dt ρ w 2dx C(ρ¯) w 2 . 2 Z | | Z k∇ kL2 ≤ 2 Z 0| 0| ≤ k 0kL2 0 Multiplying (3.15) by ∂ w and integrating over R3, one gets by Sobolev embedding that 1 t µ d ρ ∂ w 2dx+ w 2dx = (ρu )w w dx Z | t | 2 dt Z |∇ | −Z ·∇ · t (3.17) C(ρ¯) u L6 w L3 √ρ∂tw L2 ≤ k k k∇ k k k 1 ≤4 Z ρ|∂tw|2dx+Ck∇uk2L2k∇wkL2k∇wkL6 . Notice that the momentum equation can be written as ˜ µ∆w+ P = ρ∂ w (ρu )w, t − ∇ − − ·∇ where the left handside is viewed as the Helmholtz-Weyl decomposition of the right one. From this equation, µ ∆w L2 ρ∂tw L2 + (ρu )w L2. k k ≤ k k k ·∇ k INHOMOGENEOUS INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 7 It follows from Caldero´n-Zygmund inequality and the Sobolev embedding theorem that w L6 C 2w L2 k∇ k ≤ k∇ k (3.18) C ρ∂tw L2 +C (ρu )w L2 ≤ k k k ·∇ k 1 1 ≤Ckρ∂twkL2 +Ck∇ukL2k∇wkL22k∇wkL26, which, together with Young’s inequality, implies that (3.19) k∇wkL6 ≤ Ck∇2wkL2 ≤ Ckρ∂twkL2 +Ck∇uk2L2k∇wkL2. Inserting (3.19) into (3.17), we get 1 µ d (3.20) ρ ∂ w 2dx+ w 2dx C u 4 w 2 . 2 Z | t | 2dt Z |∇ | ≤ k∇ kL2k∇ kL2 By Gronwall’s inequality and the assumption A(T) 2, ≤ t (3.21) ρ ∂ w(s) 2dxds+ w(t) 2 CeC w 2 . Z Z | t | k∇ kL2 ≤ k∇ 0kL2 0 For fixed (ρ,u), the map from w to w(t) is linear. Furthermore, by Lemmas 2.1 and 0 ∇ 2.3, L4(0,T;L2) = [L2(0,T;L2), L∞(0,T;L2)]1 andH˙ 21 L2,H˙ 1 , thenonegets, upon 2,4 ⊆ h i1,4 2 combining the estimate (3.16) and (3.21), and using Lemma 2.4, that (3.22) k∇wkL4(0,T;L2) ≤ C1kw0kH˙ 21 . Notice that C does not depend on T, since one can scale [0,T] to [0,1]. Consequently, 1 k∇ukL4(0,T;L2) ≤ C1ku0kH˙ 21. Regarding the decay of u L2, upon multiplying (3.20) by t, one gets that k∇ k d t ρ ∂ u 2dx+ t u 2dx u 2dx+Ct u 4 u 2 . Z | t | dt Z |∇ | ≤ Z |∇ | k∇ kL2k∇ kL2 If A(T) 2, then by Gronwall’s inequality, ≤ T t √ρ∂ u 2 dt+ sup t u(t) 2 Z k t kL2 k∇ kL2 0 t∈[0,T] (3.23) T C u 2 dt C ρ u 2dx, ≤ Z k∇ kL2 ≤ Z 0| 0| 0 where the basic energy inequality for u is utilized. (cid:3) Proof of Theorem 3.1. The conclusion of Theorem 3.1 will follow if we let ε 1/C , which 1 ≤ (cid:3) is given in Lemma 3.3. Remark 3.1. According to the proof of Lemma 3.3, the H˙ 21-norm of u0 can be replaced by 1 B˙ 2 -norm, which will make the main result more refined. 2,4 8 WALTER CRAIG,XIANGDIHUANG,AND YUNWANG From this point on, we will consider only the small data problem, assuming that the initial data satisfies the condition ku0kH˙ 21 ≤ ε, as in Theorem 3.1. The notationC¯ is used to denote a positive constant, which may depend on T and the initial data, and it may change from line to line. The computation is standard, with respect to that in [22], but we sketch it here for completeness. From the proof of Lemma 3.3, we know that T (3.24) √ρ∂ u 2 dt+µ u 2 C u 2 C¯ . Z k t kL2 k∇ kL∞(0,T;L2) ≤ k∇ 0kL2 ≤ 0 Lemma 3.4 (Estimates for √ρ∂tu L2 and u H1). Under the assumption of Theorem 3.1, k k k∇ k we have T (3.25) sup √ρ∂ u 2 + u 2 + ∂ u 2 dt C¯ . k t kL2 k∇ kH1 Z k∇ t kL2 ≤ t∈[0,T](cid:2) (cid:3) 0 Proof. Differentiating the momentum equation with respect to t, multiplying by ∂ u, and t then integrating over R3, one can obtain that 1 d ρ ∂ u 2dx+µ ∂ u 2dx 2dt Z | t | Z |∇ t | (3.26) = 2 ρu ∂ u ∂ udx ρu (u u ∂ u)dx (ρ∂ u )u ∂ udx . − Z ·∇ t · t −Z ·∇ ·∇ · t −Z t ·∇ · t It follows from Sobolev embedding theorem and Gagliardo-Nirenberg inequality that 2 ρu ∂ u ∂ udx − Z ·∇ t · t 3 1 (3.27) ≤ Ck∇ukL2k∇∂tukL22k√ρ∂tukL22 µ ∂ u 2 +C u 4 √ρ∂ u 2 . ≤ 8k∇ t kL2 k∇ kL2k t kL2 By Sobolev embedding and the estimate (3.19) , the second term can be estimated, ρu (u u ∂ u)dx −Z ·∇ ·∇ · t ρ u u 2 ∂ u dx+ ρ u 2 2u ∂ u dx+ ρ u 2 u ∂ u dx ≤Z | ||∇ | | t | Z | | |∇ || t | Z | | |∇ ||∇ t | (3.28) ≤CkukL6k∇ukL2k∇ukL6k∂tukL6 +Ckuk2L6k∇2ukL2k∂tukL6 +Ckuk2L6k∇ukL6k∇∂tukL2 µ C u 4 2u 2 + ∂ u 2 ≤ k∇ kL2k∇ kL2 8k∇ t kL2 µ C u 4 ρ∂ u 2 +C u 10 + ∂ u 2 . ≤ k∇ kL2k t kL2 k∇ kL2 8k∇ t kL2 INHOMOGENEOUS INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 9 For the third term on the right handside of (3.26), utilizing Gagliardo-Nirenberg inequality and (3.19), −Z (ρ∂tu·∇)u·∂tudx ≤ Ck√ρ∂tukL2k∇ukL6k√ρ∂tukL3 3 1 (3.29) ≤Ck√ρ∂tukL22k∇∂tukL22 kρ∂tukL2 +k∇uk3L2 10 (cid:0) µ (cid:1) C √ρ∂ u 3 +C √ρ∂ u 2 u 4 + ∂ u 2 . ≤ k t kL2 k t kL2k∇ kL2 8k∇ t kL2 Collecting all the estimates (3.27)-(3.29), one gets that 1 d µ ρ ∂ u 2dx+ ∂ u 2dx 2dt Z | t | 2 Z |∇ t | (3.30) 4 C u 4 √ρ∂ u 2 +C √ρ∂ u 3 √ρ∂ u 2 +C u 10 . ≤ k∇ kL2k t kL2 k t kL2k t kL2 k∇ kL2 Utilizing Gronwall’s inequality and the estimate (3.24), we obtain that T (3.31) sup √ρ∂ u 2 + ∂ u 2 dt C¯. k t kL2 Z k∇ t kL2 ≤ t∈[0,T] 0 According to the estimate (3.19), (3.31) implies that (3.32) sup k∇2ukL2 ≤ C sup k√ρ∂tukL2 +Ck∇uk3L2 ≤ C¯ . t∈[0,T] t∈[0,T] (cid:3) It completes the proof. Lemma 3.5 (Estimate for ρ H1). Under the assumptions of Theorem 3.1, we know that k∇ k T (3.33) sup k∇ρkH1 +Z k∇3uk2L2dt ≤ C¯ . t∈[0,T] 0 Proof. Differentiating the mass equation with respect to x , j = 1,2,3, we find that ∂ ρ j j satisfies (3.34) ∂ ∂ ρ+u ∂ ρ = ∂ u ρ. t j j j ·∇ − ·∇ Then multiplying (3.34) by ∂ ρ, integrating over R3, and summing over the index j, one gets j that d (3.35) dt Z |∇ρ|2dx ≤ CZ |∇u|·|∇ρ|2dx ≤ Ck∇ukL∞k∇ρk2L2 . T To derive the appropriate bound for 0 k∇ukL∞dt, we make use of the elliptic estimates related with the momentum equation. IRn fact, 2u L4 C ρ∂tu L4 +C (ρu )u L4 k∇ k ≤ k k k ·∇ k 3 1 ≤ Ck∇∂tukL42k√ρ∂tukL42 +CkukL12k∇ukL6 3 1 3 5 C ∂ u 4 √ρ∂ u 4 +C u 4 u 4 , ≤ k∇ t kL2k t kL2 k∇ kL2k∇ kL6 10 WALTER CRAIG,XIANGDIHUANG,AND YUNWANG which together with Lemma 3.4 gives that (3.36) 2u L2(0,T;L4) C¯ . k∇ k ≤ By virtue of the Gagliardo-Nirenberg inequality, 1 2 k∇ukL∞ ≤ Ck∇ukL36k∇2ukL34 , ¯ hence u L1(0,T;L∞) C. Consequently, by Gronwall’s inequality, (3.35) gives a bound for k∇ k ≤ ρ L2. Similar argument shows that k∇ k d 2ρ 2dx C u 2ρ 2 + 2u ρ 2ρ dx dt Z |∇ | ≤ Z |∇ ||∇ | |∇ ||∇ ||∇ | (cid:0) (cid:1) (3.37) ≤ Ck∇ukL∞k∇2ρk2L2 +Ck∇2ukL3k∇ρkL6k∇2ρkL2 ≤ C k∇ukL∞ +k∇2ukL3 k∇2ρk2L2 . (cid:0) (cid:1) Note that k∇2ukL3 ≤ k∇2ukL1/23k∇2uk2L/43. Combining this fact with (3.32) and (3.36), we get that 2u L1(0,T;L3) . ∇ ∈ Hence, the Gronwall inequality gives a bound for 2ρ L2. k∇ k Finally, using the reguality theory for Stokes equations, one can obtain that 3u L2 C( ρ∂tu H1 + ρu u H1) k∇ k ≤ k k k ·∇ k (3.38) ≤ C(k∇ρkL3 +1) k∇∂tukL2 +k∇uk2H1 , (cid:0) (cid:1) which implies that 3u L2(0,T;L2) C¯. (cid:3) k∇ k ≤ 3.2. Proof of Theoerem 1.2. With the a priori estimates in subsection 3.1 in hand, we are prepared for the proof of Thorem 1.2. Proof. According to Theorem 1.1, there exists a T > 0 such that the Cauchy problem ∗ (1.1)-(1.2) has a unique local strong solution (ρ,u,P) on R3 (0,T ], where T depends on ∗ ∗ × kρ0kL23∩H2, k∇u0kH1 and kgkL2. We will show that this local solution extends to a global one. It follows from the integrability property of the local strong solution that A(0) = 0. Hence there exists a T (0,T ) such that (3.11) holds for T = T . Set 1 ∗ 1 ∈ T∗ = sup T (ρ,u,P) is a strong solution on R3 (0,T] and A(T) 2 . { | × ≤ } Then T∗ T > 0. 1 The cla≥im is that T∗ = , for otherwise T∗ < , which we will assume for an argument ∞ ∞ by contradiction. First, it follows from Lemmas 3.4 and 3.5 that k∇u(t)kH1 ≤ C¯(T∗) , kρ(t)kL32∩H2 ≤ C¯(T∗) , for every 0 < t < T∗ , where C¯(T∗) depends on T∗ and the initial data. Secondly, µ∆u P = ρ∂ u (ρu )u, t −∇ − ·∇

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.