ebook img

On the Domain of Analyticity and Small Scales for the Solutions of the Damped-driven 2D Navier-Stokes Equations PDF

0.23 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 On the Domain of Analyticity and Small Scales for the Solutions of the Damped-driven 2D Navier-Stokes Equations

ON THE DOMAIN OF ANALYTICITY AND SMALL SCALES FOR THE SOLUTIONS OF THE DAMPED-DRIVEN 2D NAVIER–STOKES EQUATIONS Alexei A. Ilyin1 and Edriss S. Titi2 7 0 0 2 Abstract. Weobtainalogarithmicallysharpestimateforthespace-analyticityradiusofthesolutions of the damped-driven 2D Navier–Stokes equations with periodic boundary conditions and relate this n to the small scales in this system. This system is inspired by the Stommel–Charney barotropic ocean a circulationmodel. J 9 Key words: Analyticity, Gevrey regularity, Navier–Stokes equations, dissipative length scales, 1 Stommel–Charneymodel. AMS subject classification: 35B41,35Q30,37L30. ] P A 1. Introduction . h It was shown in [15] (see also [3], [12]) that the solutions of the 2D Navier–Stokes equa- t a tions with periodic boundary conditions belong to the Gevrey class of analytic functions m (if the forcing term does). Using the Gevrey regularity approach the following estimate [ for the spatial analyticity radius for the solutions that lie on the global attractor (or are 1 near it) was obtained v 0 c|Ω|1/2 3 la ≥ G2logG, (1.1) 5 1 where G = kfk |Ω|/ν2 is the Grashof number and |Ω| = L2/γ is the area of the periodic 0 L2 domain Ω = [0,L/γ]×[0,L], γ ≤ 1. 7 0 Therefore, the Fourier coefficients uˆ are exponentially small for |k| ≫ L/l , and l k a a / h naturally forms a lower bound for the small dissipative length scale for the system (see, t a for instance, [11]). m There are other ways of estimating the dissipative small length scale for the Navier– : Stokes system, for instance, in terms of the dimension of the global attractor [1], [6], [7], v i [12], [39]. TheHausdorff andfractaldimensions oftheglobalattractorsatisfythefollowing X estimate [8] (see also [6], [39]): r a dim A ≤ c G2/3(log(1+G))1/3, c = c (γ) F 1 1 1 which has been shown in [33] (following ideas of [1]) to be logarithmically sharp. If we accept the point of view that the small length scale can be defined as follows (see [7], [12], [36], [39]) 1/2 |Ω| l ∼ , (1.2) f dim A (cid:18) F (cid:19) DATE: JANUARY 18, 2007. 1 Keldysh Institute of Applied Mathematics, Russian Academy of Sciences, Miusskaya Sq. 4, 125047 Moscow, Russia, E-mail: [email protected] 2 Department of Mathematics and Department of Mechanical and Aerospace Engineering, University of California,Irvine, California 92697,USA, E-mail: [email protected]. Also: Department of Computer Science and Applied Mathematics, Weizmann Institute of Science, P.O. Box 26, Rehovot, 76100, Israel, E-mail: [email protected] 1 2 A.A. ILYINANDE.S. TITI then up to logarithmic correction we have |Ω|1/2 l ∼ . (1.3) f G1/3 This heuristic estimate for the small length scale is probably the best one can hope for sinceitmatches, uptologarithmicterm,thephysicallyassertedestimatesfortheenstrophy dissipation length scale [30] . We also observe that the estimate (1.3) is extensive, that is, independent of the size of the spatial domain provided that its shape is fixed. Another rigorousdefinition of the small length scale can be given in terms of the number ofdeterminingmodes, nodes, orvolumeelements (see[12],[14],[16],[28]andthereferences therein). It was shown that if N is sufficiently large and N equal squares of size l tile the dn periodic spatial domain, then any collection of points (one in each square) are determining for the long time dynamics of the 2D Navier–Stokes system. The best to date estimate for N was obtained in [28]: N ≤ c G, 2 where c = c (γ) depends only on the aspect ratio γ ≤ 1. (An explicit estimate for c was 2 2 2 obtained in [26]: c (γ) = (68/(γπ))1/2.) 2 Therefore the small length scale defined in terms of the lattice of determining nodes satisfies |Ω|1/2 l ≥ c (γ)−1/2 . (1.4) dn 2 G1/2 We observe that this estimate is not extensive, that is, l scales like λ−1/2 if Ω is replaced dn by λΩ, λ > 0. Wepoint out here that for the2DNavier–Stokes system withanalytic forcing the results of [17], [18] provide the existence of a finite number N of instantaneously determining nodes comparable with the fractal dimension of the attractor. These nodes, however, can be chosen arbitrarily (up to a subset of ΩN with 2N-dimensional Lebesgue measure zero) and therefore do not naturally define a regular lattice of determining nodes. The best to date estimate for the analyticity radius of the solutions of the Navier–Stokes equations with analytic forcing term f was obtained in [31]: |Ω|1/2 l ≥ c (γ) . (1.5) a 3 G1/2(1+logG)1/4 Relating the radius of analyticity to the dissipative small length scale (see also [23] in this regard) we note that up to a logarithmic correction the estimate (1.5) coincides with (1.4), but both are worse than (1.2), where the latter coincides, as we have already pointed out, with the physically asserted estimate of [30]. In this paper we focus on the 2D space periodic Navier–Stokes system with damping 2 ∂ u+ ui∂ u = −µu+ν∆u−∇p+f, t i (1.6) i=1 X divu = 0. By adding the Coriolis forcing term to (1.6) one obtains the well-known Stommel–Charney barotropic model of ocean circulation [4], [10], [35], [37]. Here the damping µu represents the Rayleigh friction term and f is the wind stress. For an analytical study of this system see, for instance [5], [22], [24], [41], and the references therein. In a follow up work we will be studying the effect of adding rotation (Coriolis parameter) on the size of small scales and the complexity of the dynamics of (1.6). Therefore, we will focus in this work on the system (1.6). We also point out that in this geophysical context the viscosity plays a much ANALYTICITY OF THE SOLUTIONS OF THE 2D DAMPED-DRIVEN NS SYSTEM 3 smaller role in the mechanism of dissipating energy than the Rayleigh friction. That is why in this work the friction coefficient µ > 0 will be fixed and we consider the system at the limit when ν → 0+. Sharp estimates (as ν → 0) for the Hausdorff and the fractal dimensions of the global attractor of the system (1.6) were first obtained in the case of the square-shaped domain in [25] (γ = 1). Then the case of an elongated domain was studied in [27] (γ → 0), where it was shown that krotfk |Ω| ∞ dim A ≤ c D, D = , (1.7) F 4 µν where c is an absolute constant (c ≤ 12). This estimate is sharp as both ν → 0 and 4 4 γ → 0. Therefore the small length scale defined as in (1.2) is of the order of |Ω| 1/2 µν 1/2 |Ω|1/2 l ∼ ∼ ∼ . (1.8) f dim A krotfk D1/2 (cid:18) F (cid:19) (cid:18) ∞(cid:19) This heuristic estimate is, in fact, a rigorous bound for the small length scale expressed in terms of the number of determining modes and nodes [26]: 1/2 1/2 |Ω| µν l = c = c , c = 681/4. (1.9) dn 5 5 5 D krotfk (cid:18) (cid:19) (cid:18) ∞(cid:19) This means that any lattice of points in Ω at a typical distance l ≤ l is determining. dn The main result of this paper is in showing that the analyticity radius l of the solutions a of the damped-driven Navier–Stokes system (1.6) lying on the global attractor is bounded from below and satisfies the estimate: c|Ω|1/2 l ≥ , (1.10) a D1/2(1+logD)1/2 which up to a logarithmic correction agrees both with the smallest scale estimate (1.8) and the rigorously defined typical distance between the determining nodes (1.9). It is worth mentioning that this point of view of relating the radius of analyticity of solutions on the Navier–Stokes equations to small scales in turbulence was also presented in [23]. This paper is organized as follows. In section 2 we employ the Gevrey–Hilbert space technique of [15] to derive a lower bound for the radius of analyticity of the order c|Ω|1/2 . (1.11) D2logD This bound considerably improves, for a fixed µ > 0, the lower bound (1.1) for the clas- sical Navier–Stokes system as ν → 0+ (see also Remark 2.1). Let us remark that as an alternative to the Gevrey regularity technique for estimating small scales one can apply the ladder estimates approach presented in [9] to obtain estimates for the small scales in (1.6) (see also [19]). In section 3 the estimate (1.10) is proved for the system (1.6) following [31]. 2. Gevrey regularity of the damped Navier–Stokes system As usual (see, for instance, [1],[6],[32],[38]), we write (1.6) as an evolution equation in the Hilbert space H which is the closed subspace of solenoidal vectors in (L (Ω))2 with 2 zero average over the torus Ω = [0,L/γ]×[0,L]: ∂ u+B(u,u)+νAu+µu = f, u(0) = u . (2.1) t 0 4 A.A. ILYINANDE.S. TITI Here A = −P∆ is the Stokes operator with eigenvalues 0 < λ ≤ λ ≤ ..., B(u,v) = 1 2 P 2 ui∂ v is the nonlinear term, f = Pf ∈ H, and P : (L (Ω))2 → H. i=1 i 2 We restrict ourselves to the case γ = 1 and, in addition, assume that Ω = [0,2π]2 (this (cid:0)P (cid:1) simplifies the Fourier series below). The case of the square-shaped domain Ω = [0,L]2 reduces to this case by scaling. Furthermore, any domain with aspect ratio γ < 1 can be treated in the similar way, the absolute dimensionless constants c ,c ,... below will then 1 2 depend on γ, however. A vector field u ∈ H has the Fourier series expansion u = u eij·x, u ∈ C2, u = u¯ , u ·j = 0, u = 0, j j −j j j 0 j∈Z2 X and kuk2 = kuk2 = (2π)2 |u |2. L2 j j∈Z2 X The eigenvalues of the Stokes operator A are the numbers |j|2, and the domain of its powers is the set of vector functions u such that (2π)2 |j|4α|u |2 = kAαuk2 < ∞. j j∈Z2 X For τ,s > 0 we define the Gevrey space D(eτAs) of functions u satisfying (2π)2 e2τ|j|2s|u |2 = keτAsuk2 < ∞. (2.2) j j∈Z2 X We suppose that the forcing term f belongs to the Gevrey space of analytic functions f ∈ D(eσ1A1/2A1/2), so that (2π)2 |j|2e2σ1|j||u |2 = keσ1A1/2A1/2fk < ∞ (2.3) j j∈Z2 X for some σ > 0. We set 1 1/2 ϕ(t) = min(νλ t,σ ). 1 1 The norm and the scalar product in D(eϕ(t)A1/2) are denoted by k·k and (·,·) , respec- ϕ ϕ tively. Weassume thatu ∈ D(A1/2)andtake thescalar product of(2.1)andAuinD(eϕ(t)A1/2) 0 1/2 for sufficiently small t ≤ σ /(νλ ). Since 1 1 1 eϕ(t)A1/2∂ u(t),eϕ(t)A1/2u(t) = ∂ kA1/2u(t)k2 −νλ1/2(Au(t),A1/2u(t)) , t 2 t ϕ 1 ϕ (cid:0) (cid:1) we obtain 1 ∂ kA1/2uk2 +νkAuk2 +µkA1/2uk2 = (B(u,u)Au) +νλ1/2(Au,A1/2) +(A1/2f,A1/2u) . 2 t ϕ ϕ ϕ ϕ 1 ϕ ϕ (2.4) Next we use the key estimate (see [15], [12], [40]) for the nonlinear term in Gevrey spaces kAuk2 1/2 (B(u,u),Au) ≤ c kA1/2uk2kAuk 1+log ϕ ϕ 1 ϕ ϕ λ kA1/2uk2 (cid:18) 1 ϕ(cid:19) ANALYTICITY OF THE SOLUTIONS OF THE 2D DAMPED-DRIVEN NS SYSTEM 5 and use Young’s inequality for this estimate and for the last two terms in (2.4): ∂ kA1/2uk2+νkAuk2 ≤ t ϕ ϕ 2c2 kAuk2 kA1/2fk2 ≤ 1kA1/2uk4 1+log ϕ +2νλ kA1/2uk2 + ϕ ≤ ν ϕ λ kA1/2uk2 1 ϕ 2µ (cid:18) 1 ϕ(cid:19) c kAuk2 kA1/2fk2 ≤ 2kA1/2uk4 1+log ϕ +ν3λ2 + ϕ, ν ϕ λ kA1/2uk2 1 2µ (cid:18) 1 ϕ(cid:19) where c = 2c2 +1. Next, using the inequality −αz +β(1+logz) ≤ βlogβ/α (see [12], 2 1 [13]), we find c kAuk2 c c kA1/2uk2 −νkAuk2 + 2kA1/2uk4 1+log ϕ ≤ 2kA1/2uk4 log 2 ϕ, ϕ ν ϕ λ kA1/2uk2 ν ϕ λ ν2 (cid:18) 1 ϕ(cid:19) 1 and obtain the differential inequality c c kA1/2uk2 kA1/2fk2 ∂ kA1/2uk2 ≤ 2kA1/2uk4 log 2 ϕ +ν3λ2 + ϕ . t ϕ ν ϕ λ ν2 1 2µ 1 Hence the function c kA1/2uk2 kA1/2fk2 y(t) = 2 ϕ + σ1 +e, λ ν2 λ ν3/2µ1/2 1 1 where lne = 1, satisfies ∂ y(t) ≤ νλ c y2logy, c = max(1,c /2). t 1 3 3 2 Therefore y(t) ≤ 2y(0) for as long as t ≤ (2νλ c y(0)log2y(0))−1. 1 3 In other words, kA1/2uk2 ≤ 2kA1/2u k2 +c (ν/µ)1/2kA1/2fk +c λ ν2, c = e/c , ϕ 0 4 σ1 4 1 4 2 as long as 0 ≤ t ≤ T∗(kA1/2u k), where 0 1 T∗(kA1/2u k) = . 0 2c νλ c2kA1/2u0k2 + kA1/2fkσ1 +e log 2 c2kA1/2u0k2 + kA1/2fkσ1 +e 3 1 λ1ν2 λ1ν3/2µ1/2 λ1ν2 λ1ν3/2µ1/2 (cid:16) (cid:17) (cid:16) (cid:16) (cid:17)(cid:17) We now observe (see Lemma 3.1) that on the global attractor or in the absorbing ball we have, respectively, kA1/2fk kA1/2fk kA1/2u(t)k ≤ , t ∈ R, kA1/2u(t)k ≤ 2 , t ≥ T (kA1/2u k). 0 0 µ µ Therefore we have the following lower bound for T∗: kA1/2fk2 kA1/2fk kA1/2fk2 kA1/2fk −1 T∗ ≥ c νλ + σ1 +1 log + σ1 +1 5 1 λ ν2µ2 λ ν3/2µ1/2 λ ν2µ2 λ ν3/2µ1/2 (cid:20) (cid:18) 1 1 (cid:19) (cid:18) 1 1 (cid:19)(cid:21) In the limit ν → 0+ we have kA1/2fk2 kA1/2fk ≫ σ1, λ ν2µ2 λ ν3/2µ1/2 1 1 and we can write the lower bound for T∗ as follows T∗ ≥ c νλ D2logD −1, 6 1 (cid:2) (cid:3) 6 A.A. ILYINANDE.S. TITI where kA1/2fk krotfk|Ω|1/2 1 krotfk |Ω| ∞ = ≤ D, where D = . λ1/2νµ 2πνµ 2π νµ 1 In terms of the analyticity radius l the lower bound for T∗ takes the form a c |Ω|1/2 7 l ≥ . a D2logD Thus, we have proved the following theorem. Theorem 2.1. Suppose that f ∈ D(A1/2eσ1A1/2) for some σ > 0. Then a solution u lying 1 on the global attractor A is analytic with analyticity radius c |Ω|1/2 7 l ≥ min , σ , a (D2 +D +1)log(D2 +D +1) 1 (cid:18) 1 1 (cid:19) where krotfk |Ω| kA1/2fk D = ∞ , D = σ1. νµ 1 λ ν3/2µ1/2 1 Moreover, c |Ω|1/2 l ≥ 8 as ν → 0+. (2.5) a D2logD The constants c and c depend only on the aspect ratio of the periodic domain Ω. 7 8 Remark 2.1. We observe that the estimate (2.5) for the system (1.6) is of the order ν−2log(1/ν) as far as the dependence on ν → 0+ is concerned, while the estimate (1.1) for the classical Navier–Stokes system is, in this respect much larger; namely, is of the order ν−4log(1/ν). However, theestimate (2.5)isnotsharpandwillbeimproved inthenext section. Ashas been demonstrated in [34] the Gevrey–Hilbert space technique does not always provide sharp estimates for the radius of analyticity. The mechanism explaining this has been reported in [34] by means of an explicitly solvable model equation. 3. Sharper bounds In this section we obtain sharper lower bounds for the analyticity radius l . This is a achieved by combining the ν-independent estimate for the vorticity contained in the fol- lowing lemma and the L -technique developed in [21], [31] for the uniform analyticity p radius of the solutions of the Navier–Stokes equations. We observe that similar technique has been earlier established in [2] for studying the analyticity of the Euler equations. Applying the operator rot to (1.6) we obtain the well-known scalar vorticity equation ∂ ω +u·∇ω = ν∆ω −µω +F, (3.1) t where ω = rotu, F = rotf, u = ∇⊥∆−1ω, so that u·∇ω = ∇⊥∆−1ω·∇ω = J(∆−1ω,ω), and ∇⊥ = (−∂ ,∂ ), J(a,b) = ∇⊥a·∇b. 2 1 Lemma 3.1. (See [26].) The solutions u(t) lying on the global attractor A satisfy the following bound: krotfk kω(t)k ≤ L2k , t ∈ R, (3.2) L2k µ where 1 ≤ k ≤ ∞. ANALYTICITY OF THE SOLUTIONS OF THE 2D DAMPED-DRIVEN NS SYSTEM 7 Proof. We use the vorticity equation (3.1) and take the scalar product with ω2k−1, where k ≥ 1 is integer, and use the identity (J(ψ,ϕ),ϕ2k−1) = (2k)−1 J(ψ,ϕ2k)dx = (2k)−1 div(ϕ2k∇⊥ψ)dx = 0. Z Z We obtain kωk2k−1∂ kωk +(2k −1)ν |∇ω|2ω2k−2dx+µkωk2k = L2k t L2k L2k Z = (rotf,ω2k−1) ≤ krotfk kωk2k−1. L2k L2k Hence, by Gronwall’s inequality kω(t)k ≤ kω(0)k e−µt +µ−1krotfk (1−e−µt), L2k L2k L2k and passing to the limit as k → ∞ we find kω(t)k ≤ kω(0)k e−µt +µ−1krotfk (1−e−µt). ∞ ∞ ∞ Now, we let t → ∞ in the above inequalities and obtain krotfk limsupkω(t)k ≤ L2k , 1 ≤ k ≤ ∞, L2k µ t→∞ which gives (3.2) since the solutions lying on the attractor are bounded for t ∈ R. (cid:3) As before we consider the square-shaped domain Ω = [0,L]2 and it is now convenient to write (1.6) in dimensionless form. We introduce dimensionless variables x′, t′, u′ and p′ by setting x = Lx′, t = (L2/ν)t′, u = (ν/L)u′, p = (ν2/L2)p′, µ = (ν/L2)µ′. We obtain 2 ∂t′u′ + u′i∂i′u′ = −µ′u+∆′u′ −∇′p′ +f′, (3.3) i=1 X div′u′ = 0, where x′ ∈ Ω′ = [0,1]2, f′ = (L3/ν2)f. Accordingly, the dimensionless form of (3.1) is as follows (we omit the primes): ∂ ω +u·∇ω = ∆ω −µω +F. (3.4) t Remark 3.1. For dimensionless variables u′ and ω′ the estimate (3.2) with k = ∞ takes the form krot′f′k krotfk L2 kω′k = krot′u′k ≤ ∞ = ∞ = D. (3.5) ∞ ∞ µ′ νµ The next lemma is similar to the main estimate for the space analyticity radius in [31]. Lemma 3.2. Suppose that F is a restriction to Ω (that is, y = 0) of a bounded x-periodic analytic function F(x+iy)+iG(x+iy) in the region |y| ≤ δ and F M2 = sup (F(x+iy)2 +G(x+iy)2). (3.6) F x∈Ω, |y|≤δF Let p ≥ 3/2 and let M2 2p t = . 0 CM2/µ F 8 A.A. ILYINANDE.S. TITI Here (and throughout) C is a sufficiently large universal constant and M ≥ kω k . 2p 0 L2p Then the solution ω(t) is analytic for t > 0 and for 0 < t ≤ t the space analyticity radius 0 of ω(t) is greater than t1/2 1 1 1 δ(t) = min , , , ,δ . C Cpt(2p−3)/4pM2p Cpt(2p−3)/(4p+6)M2p/(2p+3) pt1/2M2p F! 2p Proof. We solve (3.4) by a sequence of approximating solutions (see [29], [31]). We set u(0) = 0 and ω(0) = 0. Then for ω(n), u(n) we have the equation ∂ ω(n) −∆ω(n) +u(n−1) ·∇ω(n) +µω(n) = F t (3.7) ω(n)(0) = ω = rotu , u(n) = ∇⊥∆−1ω(n). 0 0 The solutions ω(n) and u(n) for t > 0 have analytic extensions ω(n)+iθ(n) and u(n)+iv(n) and since the system (3.7) is linear, their analyticity radius is at least δ . They satisfy F the equation ∂ (ω(n)+iθ(n))−∆(ω(n)+iθ(n))+(u(n−1)+iv(n−1))·∇(ω(n)+iθ(n))+µ(ω(n)+iθ(n)) = F+iG, t or, equivalently, the system ∂ ω(n) −∆ω(n) +µω(n) +u(n−1) ·∇ω(n) −v(n−1) ·∇θ(n) = F, t (3.8) ∂ θ(n) −∆θ(n) +µθ(n) +u(n−1) ·∇θ(n) +v(n−1) ·∇ω(n) = G, t where, as before, u(n) = ∇⊥∆−1ω(n), v(n) = ∇⊥∆−1θ(n), and the differential operators are taken with respect to x. In view of the analyticity of the solutions we have the Cauchy– Riemann equations ∂ω(n) ∂θ(n) = − , ∂y ∂x j j (3.9) ∂ω(n) ∂θ(n) = , j = 1,2, ∂x ∂y j j and the similar equations for u(n) and v(n). Let ε > 0. We consider the functional 1 ψ (t) = ω(n)(x,αts,t)2 +θ(n)(x,αts,t)2 +ε pdxds. (3.10) n Z0 ZΩ (cid:0) (cid:1) We also set Q (x,s,t) = ω(n)(x,αts,t)2 +θ(n)(x,αts,t)2 +ε. n Here t ∈ R+ and α ∈ R2. The combination αts will play the roleof the variable y; p ≥ 3/2, and ε > 0 is arbitrary. We differentiate ψ (t) taking into account (3.8) and use the Cauchy–Riemann equa- n tions (3.9) to handle the derivatives with respect to y. We obtain 1 ∂ ψ (t)+I = I +I +I +I , (3.11) t n 0 1 2 3 4 2p where 1 I = Qp−1 |∇ω(n)|2 +|∇θ(n)|2 +µ(ω(n))2 +µ(θ(n))2 dxds+ 0 n Z0 ZΩ (cid:0) 1 (cid:1) +2(p−1) Qp−2 ω(n)∇ω(n) +θ(n)∇θ(n) 2dxds, n Z0 ZΩ (cid:0) (cid:1) ANALYTICITY OF THE SOLUTIONS OF THE 2D DAMPED-DRIVEN NS SYSTEM 9 and 1 I = Qp−1 −ω(n)∇θ(n) +θ(n)∇ω(n) ·αsdxds, 1 n Z0 ZΩ 1 (cid:0) (cid:1) I = Qp−1 ω(n)∇ω(n) +θ(n)∇θ(n) ·u(n−1)dxds, 2 n Z0 ZΩ 1 (cid:0) (cid:1) I = Qp−1 −ω(n)∇θ(n) +θ(n)∇ω(n) ·v(n−1)dxds, 3 n Z0 ZΩ 1 (cid:0) (cid:1) I = Qp−1 ω(n)F +θ(n)G)dxds. 4 n Z0 ZΩ The arguments of Q are x,s,t, an(cid:0)d the arguments of ω(n), θ(n), u(n), and v(n) are x, αts, n and t. For an arbitrary η > 0 we have 1 I ≤ η Qp−1 |∇ω(n)|2 +|∇θ(n)|2 dxds+ 1 n Z0 ZΩ (3.12) 1 (cid:0) (cid:1) C Qp−1 (ω(n))2 +(θ(n))2 |α|2s2dxds ≤ ηI +C |α|2ψ (t). η n 0 η n Z0 ZΩ (cid:0) (cid:1) Next, 1 1 I = ∇Qp ·u(n−1)dxds = 0. (3.13) 2 2p n Z0 ZΩ For I we have 3 1 I ≤ ηI +C Qp |v(n−1)|2dxds ≤ ηI +C I′I′′, (3.14) 3 0 η n 0 η 3 3 Z0 ZΩ where 1 (p−1)/p 2 1 1/p I′ = Q (x,s,t)p2/(p−1)dxds , I′′ = |v(n−1)(x,αts,t)|2pdxds . 3 n 3 j (cid:18)Z0 ZΩ (cid:19) j=1 (cid:18)Z0 ZΩ (cid:19) X We write I′ as follows 3 I′ = kQp/2k2 , Ω = Ω×[0,1] ⊂ R3, β = 2p/(p−1), 2 ≤ β ≤ 6, 3 n Lβ(Ω0) 0 and use in Ω the Gagliardo–Nirenberg inequality 0 3/β−1/2 3/2−3/β kAk ≤ CkAk k∇ Ak +CkAk Lβ(Ω0) L2(Ω0) x,s L2(Ω0) L2(Ω0) p/2 for A = A(x,s) = Q (x,s,t). We have n k∇ Ak2 = k∇ Qp/2k2 = x,s L2(Ω0) x,s n L2(Ω0) 1 p2 Qp−2 (ω(n)∇ω(n) +θ(n)∇θ(n))2 +t2(θ(n)α·∇ω(n) −ω(n)α·∇θ(n))2 dxds ≤ n Z0 ZΩ ≤ C(cid:0)p2(1+|α|2t2)I . (cid:1) 0 Hence, k∇ Ak3/2−3/β = k∇ Qp/2k3/2p ≤ C(1+|α|2t2)3/4pI3/4p. x,s L2(Ω0) x,s n L2(Ω0) 0 Next, kQp/2k2 = ψ (t), n L2(Ω0) n kAk3/β−1/2 = kQp/2k(2p−3)/2p = ψ (t)(2p−3)/4p L2(Ω0) n L2(Ω0) n 10 A.A. ILYINANDE.S. TITI and I′ ≤ C(1+|α|2t2)3/2pI3/2pψ (t)(2p−3)/2p +Cψ (t). (3.15) 3 0 n n We now consider I′′. Since v(n−1)(x,0,t) = 0 (the solution restricted to y = 0 is real- 3 j valued), we have (using the Cauchy–Riemann equations for v ) j 2 1 (n−1) (n−1) |v (x,αts,t)| = α ts ∂ v (x,αtsτ,t)dτ = j k yk j (cid:12)(cid:12)Xk=1 Z0 (cid:12)(cid:12) (cid:12) (cid:12) (cid:12) 2 1 (cid:12) (cid:12) α ts ∂ u(n−1)(x,αtsτ,t)dτ (cid:12). k k j (cid:12)(cid:12)Xk=1 Z0 (cid:12)(cid:12) (cid:12) (cid:12) Then (cid:12) (cid:12) (cid:12) (cid:12) 1/p 2p 2 1 2 1 I′′ = α ts ∂ u(n−1)(x,αtsτ,t)dτ dxds ≤ 3  k k j  Xj=1 Z0 ZΩ(cid:12)(cid:12)Xk=1 Z0 (cid:12)(cid:12) (cid:12) (cid:12)  (cid:12) 1 1 (cid:12) 1/p C|α|2t2 (cid:12) ∇u(n−1)(x,αtsτ,t) 2ps2p(cid:12)dτdxds = (cid:18)Z0 ZΩZ0 (cid:19) C|α|2t2 1s2pds (cid:12)(cid:12) 1dτ ∇u(n−1)(x(cid:12)(cid:12),αtsτ,t) 2pdx 1/p. (cid:18)Z0 Z0 ZΩ (cid:19) (cid:12) (cid:12) Since u = ∇⊥∆−1ω, we have (see [20], [42]) (cid:12) (cid:12) 1/2p |∇u(x)|2pdx = k∇∇⊥∆−1ωk ≤ k∆−1ωk ≤ Cpkωk . L2p W22p L2p (cid:18)ZΩ (cid:19) Therefore 1 1 1/p I′′ ≤ Cp2|α|2t2 ω(n−1)(x,αtsτ,t) 2pdxs2pdsdτ ≤ 3 (cid:18)Z0 Z0 ZΩ (cid:19) 1 (cid:12)(cid:12) 1/p (cid:12)(cid:12) Cp2|α|2t2 Qp dxds ≤ Cp2|α|2t2ψ (t)1/p, n−1 n−1 (cid:18)Z0 ZΩ (cid:19) where we have used 1 1h(sτ)s2pdsdτ ≤ (2p)−1 1h(s)ds. Combining this with (3.14) 0 0 0 and (3.15) we obtain R R R I ≤ 3 η′I0 +Cη′p2|α|2t2(1+|α|2t2)3/2pI03/2pψn(t)(2p−3)/2pψn−1(t)1/p +Cp2|α|2t2ψn−1(t)1/pψn(t) ≤ ηI +C p2(|α|t)4p/(2p−3)(1+|α|2t2)3/(2p−3)ψ (t)2/(2p−3)ψ (t)+Cp2|α|2t2ψ (t)1/pψ (t). 0 η n−1 n n−1 n (3.16) Finally, we estimate I : 4 1 I ≤ Qp−1 (ω(n))2 +θ(n))2)ηµ+(F2 +G2)/(4ηµ) dxds ≤ 4 n Z0 ZΩ (3.17) (cid:0) 1 (cid:1) ηI +C (M2/µ) Qp−1dxds ≤ ηI +C (M2/µ)ψ (t)(p−1)/p, 0 η F n 0 η F n Z0 ZΩ where M is defined in (3.6). F

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.