• Non ci sono risultati.

PHASE SPACE ANALYSIS APPLIED TO GEOPHYSICAL FLUIDS AND THERMOELASTICITY

N/A
N/A
Protected

Academic year: 2021

Condividi "PHASE SPACE ANALYSIS APPLIED TO GEOPHYSICAL FLUIDS AND THERMOELASTICITY"

Copied!
99
0
0

Testo completo

(1)

UNIVERSIT ` A DEGLI STUDI DI TRIESTE

XXIV CICLO DEL DOTTORATO DI RICERCA IN

ENVIRONMENTAL AND INDUSTRIAL

FLUID MECHANICS

PHASE SPACE ANALYSIS APPLIED

TO GEOPHYSICAL FLUIDS AND

THERMOELASTICITY

Settore scientifico-disciplinare: MAT/05 ANALISI MATEMATICA

DOTTORANDO

MARCO PIVETTA

COORDINATORE

PROF. VINCENZO ARMENIO

SUPERVISORE

PROF. DANIELE DEL SANTO

ANNO ACCADEMICO 2011/2012

(2)
(3)

Introduction

Since it was used for the first time, the phase space analysis has become one of the most important tool in dealing with partial differential equations. The starting idea behind this theory is that there is a strong connection between space and frequency when one has to consider physical models and for this reason also mathematical analysis must take into account both variables. Fourier trasform is indeed the main tool in this context, since it is the bridge between the two visions.

One of the theories that arose from this approach and that will be widely used in this thesis thanks to its flexibility is Littlewood-Paley ([LP]) decomposition, that is based on a dyadic partition of unity in the frequency space (see e.g. [AG] or [M]).

first of all this theory allows to define Sobolev spaces in an equivalent way using the properties of the L2 norms of the dyadic blocks, then it also gives an easy way to define interpolation spaces, such as Besov spaces, that will be repeatedly used in this work. This technique gives also origin to other fundamental instruments, like Bony’s paraproduct that will be crucial to prove the inequalities necessary for one of the main results (see [B]).

The thesis is devoted to prove three theorems, the first two deals with existence and uniqueness of mild and weak solutions of Navier-Stokes equations for geophys- ical incompressible fluids, while the third concerns uniqueness for thermoelastic system. Here we give an overview of the three results.

We start from the topic discussed in chapter 2 and 3, namely Navier-Stokes equa- tions for geophysical fluids, that we will refer to as Navier-Stokes-Coriolis equa- tions, which are represented by the following system

tu − ν∆u + (u · ∇)u + Ωe3 × u + ∇p = 0

∇ · u = 0 u(0) = u0.

(1)

We focus our attention on mild solutions for this system, whose existence and uniqueness will be proved by using a suitable fix point theorem on the following

iii

(4)

equation, that can be obtained from system (1)

T u = G(t)u0+ B(u, u), (2)

with

B(u, v) := − Z t

0

G(t − τ )P∇ · (u ⊗ v)dτ,

and where G(t) is the following time dependant operator

G(t)w = F−1



cos

 Ωtξ3

|ξ|



I + sin

 Ωtξ3

|ξ|

 R(ξ)



e−νt|ξ|2w(ξ)b

 ,

See chapter 1 for precise definitions. Our purpose is to deal with an initial datum belonging to a low regular space: the idea comes from the result obtained in a similar framework on classical Navier-Stokes equations. Let us briefly describe this point. It is known that Navier-Stokes equations have a scaling property, which means that if u(t, x) is a solution of N-S system then also

uλ(t, x) = λu(λ2t, λx)

still represents a solution. Since a fixed point theorem will be used to prove exis- tence of solutions, the norm of the spaces in which this theorem will be performed must be invariant with respect to this scaling (scaling invariance is also connected with other properties of the solutions like selfsimilarity, remarked in the pioneering work [L]).

The first invariant space used in studying mild solution for three dimensional NS equations is the Sobolev space ˙H12 and this space was used by Fujita and Kato ([FK]) in 1964, while some years later Kato ([K]) generalized the previous result considering the space L3.

Thanks to the characterization of Besov spaces, Cannone, Meyer and Planchon were able to obtain existence and uniqueness result for the space ˙B

3 p−1

p,∞, providing a result in which highly oscillating initial data can be considered, since negative index of Besov space is allowed (see [CMP] and [CP]).

The final step was obtained by Koch and Tataru, who proved the sharpest result on this field ([KT]). They were able to enlarge the result in [CMP] to the space

∇BM O. The last space that could be considered is the largest scaling invariant

(5)

v space in R3, that is ˙B∞,∞−1 , but for this space a counterexample of non existence has been proved.

We can suggestively resume all these results in the following picture, which prop- erly underline the continuous imbeddings that link these spaces.

12(R3)

| {z }

F ujita−Kato

,→ L3(R3)

| {z }

Kato

,→ ˙B

3 p−1

∞,p(R3)

| {z }

Cannone M eyer P lanchon

,→ ∇BM O(R3)

| {z }

Koch−T ataru

,→ B∞,∞−1 (R3).

Before describing the technique of the proof, let us consider what can be said so far about Navier-Stokes-Coriolis equations: from [HS] we know that existence and uniqueness of mild solutions are known when the initial datum belongs to the Sobolev space ˙H12. In chapter 2 we will obtain this result using the same technique used by Chemin in [C]. In the following chapter we try to follow the road represented above by proving two theorems, the first using L3 space, the second using the Besov space ˙B

3 p−1

p,∞, but limiting ourselves to 3 < p < 4. We state here only the theorem concerning Besov space, since the first is analougous.

Theorem 1 Let 3 < p < 4. There exists a constant c > 0, depending on Ω/ν, such that if ku0kB

H˙ 1 2 , ˙B

3 p,∞p −1

≤ c, then there exist a unique fixed point u(t) ∈ K4 of (2).

Moreover it holds that

i) ku(t)kK4 ≤ 2kG(t)u0kK4, ii) u ∈ L(0, +∞; B

H˙12, ˙B

3p −1 p,∞

),

iii) B(u, u) ∈ C(0, +∞; ˙H12) and lim

t→0+kB(u(t), u(t))k˙

H12 = 0.

The space Kp is defined in this way Kp :=

(

f ∈ C(]0, +∞[; Lp)/kf kKp := sup

t∈[0,+∞[

t12(1−3p)kf kLp < +∞

) ,

while the definition of the space B

H˙12, ˙B

3p −1 p,∞

is the following:

let χB(0,1) be the characteristic function of the ball B(0, 1), we define f(1)(t, x) = F−1B(0,1)f )b

f(2)(t, x) = F−1((1 − χB(0,1)) bf ),

(6)

we say that f ∈ B

H˙12, ˙B

3p −1 p,∞

if

kf kB

H˙1 2 , ˙B

3p −1 p,∞

= kf(1)k˙

H12 + kf(2)k

B˙

3 p,∞p −1

< +∞,

in the three dimensional case it holds that

12 ,→ B

H˙12, ˙B

3p −1 p,∞

,

so this theorem improves the result in [HS], moreover the choice of Besov spaces with negative index allows us to choose highly oscillating initial data. The idea of hybrid spaces is due to the work in [CMZ], where the authors obtain a result similar to ours, considering a Ω dependent space for initial datum.

The proof of this theorem consists essentially of three steps:

i) we first study the mapping property of G(t) acting on the initial datum, mainly focusing on its codomine;

ii) we prove a bilinear estimate for B(·, ·);

iii) we make use of a fix point theorem, that gives us the existence and uniqueness of a mild solution.

In chapter 4 we will briefly address the following anisotropic Navier-Stokes-Coriolis system

tu − νhhu + ∇ · (u ⊗ u) + B(t, x1, x2) × u +1ρ∇p = 0

∇ · u = 0 u(0) = u0

(AN SC) (3) where the operator ∆h = ∂x21+ ∂x22 is the horizontal laplacian. The system is called anisotropic since 0 vertical viscosity is considered. For this system we study here only existence: in fact we prove the following theorem

Theorem 2 Let u0 ∈ B0,12 a divergence free vector fields. There exists c > 0 such that if ku0k

B0, 12 ≤ cνh then there exists a global solution u of ANSC such that u ∈ L(R+; B0,12) and ∇hu ∈ L2(R+; B0,12).

(7)

vii This theorem is the consequence of two results obtained in [MP] and [P]: in the first the system (ANSC) is studied in the anisotropic Sobolev space H0,s, for s > 12, while in the second anisotropic Navier-Stokes system (without Coriolis term) is studied in the anisotropic Besov space B0,12, since the case s = 12 can only be treated using this space instead of Sobolev one. This Besov space is defined as follows

kukB0, 12 =X

q∈Z

2q2k∆vqukL2 < +∞,

where ∆vq is the vertical Littlewood-Paley decomposition. This result does not represent a proper generalization of [MP] since

B0,12 ,→ H0,12, but it allows to go beyond the assumption s > 12.

The proof relies on Ascoli-Arzel`a theorem: first we will write a sequence of ap- proximating systems of equations and build a sequence of solutions, then we will prove that there exists a subsequence converging to the solution of (3). Littlewood- Paley decomposition will play a crucial role in different points in this context, since it will be used to prove the boundness condition necessary to use Ascoli-Arzel`a theorem.

The theorems in [MP] and [P] also deal with continuity and uniqueness of the solution. In our work we do not describe completely these two points, but we think that one can obtain a better result than the one above, since Coriolis term does not add any real difficulty and the proof in [P] can be repeated also in this case. So the theorem one could prove is the following:

Theorem 3 Let u0 ∈ B0,12 a divergence free vector fields. There exists c > 0 such that if ku0k

B0, 12 ≤ cνh then there exists a unique global solution u of ANSC such that

u ∈ Cb(R+; B0,12) and ∇hu ∈ L2(R+; B0,12).

(8)

The third result, shown in chapter 5, concerns the uniqueness of the solution for the following backward thermoelastic system





t2u − ∂x(a(t, x)∂xu) + α(t, x)∂xθ(t, x) + β(t, x)θ(t, x) = f (t, x)

tθ + ∂x(b(t, x)∂xθ) + δ(t, x)∂xtu(t, x) + ρ(t, x)∂tu(t, x) + φ(t, x)u(t, x) = g(t, x) )

]0, T [×R u(0, x) = u0(x), ∂tu(0, x) = u1(x), θ(0, x) = θ0(x) on R,

(4) where low regularity assumptions for the coefficients a(t, x) and b(t, x) are consid- ered, i.e.

a(t, x) ∈ LL(R+t , LL(Rx))

b(t, x) ∈ Cµ(R+t , C1,(Rx)), (5) where LL refers to log-lipschitz space, while µ is a modulus of continuity satistfying Osgood condition, namely

Z 1 0

1

µ(t)dt = +∞,

and another technical condition, that we will refer to as ? condition. These results improves the one by Koch and Lasiecka (see [KL]), where the same system was studied considering Lipschitz regularity in space and time for a and b. The com- bination of two Carleman inequalities, that we will describe below, used in [KL] is actually the same in this work. At the end of the chapter we will also introduce a non Lipschitz function space, whose modulus of continuity satisfies Osgood and ? condition. For a description of the use of Carleman inequalities see [Ca].

The system (4) is represented by two partial differential equations coupled: the coupling is represented by the term δ(t, x)∂xtu(t, x) that represents the influence of the motion on the heat tranfer process (see for example [Bi] and [ChS]).

The principal part of these two equations are composed by a hyperbolic and a backward parabolic operator, writen in divergence form. Studying the uniqueness problem for these operators when low regular coefficients are considered is not a new issue, but it is just a new episode of a long story. The first and pioneering result on this topic is contained in the work by Colombini, De Giorgi and Spag- nolo [CDGS] in 1979, where they first addressed the problem of uniqueness for hyperbolic operator with time dependent non-Lipschitz coefficient.

Being the source of a very fertile production of results, this work soon became a classical result, since even more than thirty years later some of the techniques used in [CDGS] are still performed in the result proved nowadays. All these results are

(9)

ix based on an energy estimate for the hyperbolic operator, with loss of derivatives.

Results on hyperbolic operators have been obtained by Tarama in [T], Colombini and Lerner [CL] and Colombini and Del Santo [CDS], just to give a (very) short list. The result in [CL] is in fact the one that inspired the result, since the regularity used here is just the same, but since we need a Carleman estimate for the energy, we also need to borrow some ideas from [CDS], where Log-Zygmund coefficients are considered.

For the backward parabolic operator our guide is the work by Del Santo and Prizzi ([DSP]) where a Carleman estimates for the backward parabolic operator is ob- tained when the coefficients are continuous in time with a modulus of continuity that satisfies Osgood condition. Due to the final combination of the two inequali- ties we will use in this work, we need one more hypotesis, the so called ? condition, that we will describe precisely in the first chapter.

Since the system is linear in order to prove uniqueness it is sufficient to prove that the homogeneous system associated with the one above has only the trivial solution. In order to do this we first obtain two Carleman estimates involving the operators Pu = ∂t2u − ∂x(a(t, x)∂xu) and Lθ = ∂tθ + ∂x(b(t, x)∂xθ) and then we combine this two inequalities to conclude the proof. The weights used in the Carleman estimates are defined using the properties of the modulus of continuity µ, that will be descibed in chapter 1.

The two inequalities are the following

Z T2

0

e2γΦ(γ(T −t))E(t)dt ≤ C0 Φ0 γT2 2

Z T2

0

e2γΦ(γ(T −t))kPuk2H−ω−β∗tdt

Z T2

0

eγ2Φ(γ(T −t))kLθk2H−1−ω−β∗tdt ≥

≥ K Z T2

0

e2γΦ(γ(T −t))



γkθk2H−1−ω−β∗t +√

γkθk2H−ω−β∗t + 1

0(γT ))kθk2H1−ω−β∗t

 dt,

where the energy E(t) is equivalent to k∂tukH−ω−β∗t+kukH−ω−β∗t+kukH1−ω−β∗t. Φ(·) is a function defined using µ, while ω belongs to ]0,12[. The two inequalities hold for any γ greater than a fixed γ0. As stated above for the proof of these two inequalities a very important role will be played by Bony’s paraproduct, necessary to address difficulties arising with commutators.

(10)

The combination of the two inequalities allows us to prove that E(t) ≡ 0 for t beloging to a small time interval, this implies the following theorem

Theorem 4 Let

v ∈ H2([0, T ], L2(Rx)) ∩ H1([0, T ], H1(Rx)) ∩ L2([0, T ], H2(Rx)) ζ ∈ H1([0, T ], L2(R)) ∩ L2([0, T ], H2(Rx))

solutions of (4), with f ≡ g ≡ u0 ≡ u1 ≡ θ0 ≡ 0.

Then under the hypotesis (5) we have that v ≡ ζ ≡ 0.

(11)

Contents

Introduction iii

1 Background 1

1.1 Navier-Stokes-Coriolis equations . . . 1

1.1.1 Semigroup computation . . . 2

1.1.2 Mild solution . . . 4

1.1.3 Existence of mild solutions . . . 6

1.2 Littlewood-Paley theory . . . 7

1.3 Modulus of continuity . . . 12

1.4 Functional spaces . . . 14

2 Mild solutions for the Navier-Stokes-Coriolis system: the ˙H12 case 17 2.1 Technical results . . . 18

2.2 Main theorem . . . 25

3 Mild solutions for the Navier-Stokes-Coriolis system: the L3 and Besov case 29 3.1 The ˙H12 − L3 case . . . 29

3.2 The case ˙H12 − ˙B 3 p−1 p,∞ . . . 42

4 Anisotropic Navier-Stokes-Coriolis system 45 4.1 Main Theorem . . . 46

5 Backward uniqueness for thermoelastic system 53 5.1 Carleman estimate for the hyperbolic operator . . . 54

5.2 Carleman Estimate for parabolic operator . . . 58

5.3 Statement and Proof of the main theorem . . . 69

5.4 LogK-Lipschitz functions . . . 75

Appendix 79

xi

(12)
(13)

Chapter 1

Background

1.1 Navier-Stokes-Coriolis equations

In this work we will study Navier-Stokes equations for geophysical fluids, namely the following system:

tu − ν∆u + (u · ∇)u + Ωe3 × u + ∇p = 0

∇ · u = 0 u(0) = u0,

(1.1)

where u is the velocity field, p the pressure, ν the viscosity and Ω the Coriolis parameter, which is proportional to the inverse of Rossby number.

This system models the behaviour of a fluid in the presence of the Coriolis force, which appears when one considers the dynamics in a rotating framework: the term Ωe3× u represents the consequence of rotation on the velocity field u. For a detailed description of this system and all its features we refer to [CB].

Using the divergence free condition we can recast the system in the following way

tu − ν∆u + ∇ · (u ⊗ u) + Ωe3× u + ∇p = 0

∇ · u = 0 u(0) = u0,

(1.2)

(14)

1.1.1 Semigroup computation

We start solving, in a formal way, the following linearized system.

tu − ∆u + Ωe3× u + ∇p = 0

∇ · u = 0 u(0) = u0,

(1.3)

where we set ν = 1 to fix ideas.

For convenience we now write this system component by component:













tu1− ∆u1− Ωu2+ ∂xp = 0

tu2− ∆u2+ Ωu1 + ∂yp = 0

tu3− ∆u3 + ∂zp = 0

xu1+ ∂yu2+ ∂zu3 = 0 u(0) = u0.

(1.4)

We want to use Fourier transform, so we need the Fourier transform of p. To find it we take the divergence of the first three equation and we make use of the divergence free condition, thus obtaining:

∆p = Ω(∂xu2− ∂yu1) = Ωω3,

in which we obtain the third component ω3 of the vorticity vector field. Using Fourier transform we get:

p = −b Ωcω3

|ξ|2,

with cω3 = −i(ξ1ub2 − ξ2ub1). We write the system in frequency space and we obtain

















tub1 + |ξ|2ub1− Ωub2 + iξ1Ωcω3

|ξ|2 = 0

tub2 + |ξ|2ub2+ Ωub1+ iξ2Ωcω3

|ξ|2 = 0

tub3 + |ξ|2ub3 + iξ3Ωcω3

|ξ|2 = 0 ξ1ub1+ ξ2ub2+ ξ3ub3 = 0.

(1.5)

(15)

1.1. NAVIER-STOKES-CORIOLIS EQUATIONS 3 We multiply the first equation by iξ2, the second by iξ1 and then we subtract the second from the first, obtaining

t3+ |ξ|23+ iΩξ3ub3 = 0, and so

ub3 = i∂t3+ |ξ|2ωc3

Ωξ3 , (1.6)

substituting on the third equation we obtain:

t(i∂t3+ |ξ|23

Ωξ3 ) + |ξ|2i∂t3+ |ξ|23

Ωξ3 +iΩξ3

|ξ|23 = 0.

Or equivalently:

(∂t+ |ξ|2)23 Ωξ3 + ξ3

|ξ|2Ωcω3 = 0

Multiplying by Ωξ3 and by e|ξ|2t and calling f = ie|ξ|2t3 we can write f00+ Ω2 ξ32

|ξ|2f = 0, Writing explicitly the solution, we obtain

f (t) = A cos ξ3

|ξ|Ωt



+ B sin ξ3

|ξ|Ωt



and imposing the initial conditions we obtain f (t) = (ξ1ub2(0) − ξ2ub1(0)) cos ξ3

|ξ|Ωt



+ |ξ|ub3(0) sin ξ3

|ξ|Ωt

 .

Returning to cω3(t) we find

3(t) = −ie−|ξ|2t



1ub2(0) − ξ2ub1(0)) cos ξ3

|ξ|Ωt



+ |ξ|ub3(0) sin ξ3

|ξ|Ωt



.

Returning now in (1.6) we obtain the expression for ub3:

ub3 = e−|ξ|2t



ub3(0) cos ξ3

|ξ|Ωt



+ ξ2

|ξ| bu1(0) − ξ1

|ξ| bu2(0)



sin ξ3

|ξ|Ωt



.

(16)

Similarly we can obtain the other components:

bu(t, ξ) = cos ξ3

|ξ|Ωt



e−|ξ|2tIub0(ξ) + sin ξ3

|ξ|Ωt



e−|ξ|2tR(ξ)ub0(ξ), (1.7)

where I s the identity matrix and

R(ξ) = 1

|ξ|

0 ξ3 − ξ2

−ξ3 0 ξ1

ξ2 − ξ1 0

.

The Stokes-Coriolis semigroup can be explicitly represented by:

G(t)f = F−1



cos ξ3

|ξ|Ωt



I + sin ξ3

|ξ|Ωt

 R



eνt|ξ|2fb

 ,

1.1.2 Mild solution

Our goal is to study mild solution for Navier-Stokes-Coriolis, namely the fixed point of the map:

T u = G(t)u0+ B(u, u), with

B(u, v) := − Z t

0

G(t − τ )P∇ · (u ⊗ v)dτ,

where P is the Leray projector on divergence free vector field and G(t) is the semigroup associated with the linearized Stokes - Coriolis system.

We start working on G(t) defined by the linear problem

tu − ν∆u + Ωe3× u = f − ∇p

∇ · u = 0 u(0) = u0.

(1.8)

In the last section we obtained and wrote G(t) as

G(t)w = F−1



cos

 Ωtξ3

|ξ|



I + sin

 Ωtξ3

|ξ|

 R(ξ)



e−νt|ξ|2w(ξ)b

 .

(17)

1.1. NAVIER-STOKES-CORIOLIS EQUATIONS 5 We obtain that

u(t) = G(t)u0− Z t

0

G(t − s)(P(f(s)))ds

is solution of the problem (1.8), since G(t)u0 is the solution of

 ∂tu − ν∆u + Ωe3× u = 0

u(0) = u0, (1.9)

while Duhamel’s principle tells us that Z t

0

G(t − s)(P(f(s)))ds is the solution of

 ∂tu − ν∆u + Ωe3× u = Pf

u(0) = 0, (1.10)

Now let us compute the divergence of u(t), since ∇ · u0 = 0 and ∇ · (Pw) = 0 we get

∇ · u(t) = ∇ ·



G(t)u0+ Z t

0

G(t − s)(P(f(s)))ds



= G(t)(∇ · u0) + Z t

0

G(t − s)(∇ · (P(f(s))))ds

= 0, so ∇ · u = 0 as we want.

Now, since

P(∂tu − ν∆u + Ωe3× u) = ∂tu − ν∆u + Ωe3× u, we obtain that

P(∂tu − ν∆u + Ωe3× u − f ) = 0, and so

tu − ν∆u + Ωe3× u = Pf − ∇p.

(18)

As a consequence of these computations the solutions of the problem (1.2) will be seen as the fixed point of

u(t) = G(t)u0− Z t

0

G(t − s)(P(∇ · (u ⊗ u)))(s)ds, (1.11) with ∇ · u0 = 0.

1.1.3 Existence of mild solutions

The strategy to show existence of mild solution is the following:

• choose a space X(R3) for the initial data u0;

• choose a space Y ([0, T ], R3) such that G(t)u0 ∈ Y ;

• prove that B : Y × Y → Y is a bilinear bounded operator;

• use a fixed point argument.

The last point is represented by the following lemma:

Lemma 1 Let (Y, k · k) be a Banach space and B : Y × Y → Y a bilinear operator, such that

kB(y1, y2)k ≤ ηky1kky2k

then for any y0 ∈ Y such that 4ηky0k < 1 the equation y = y0 + B(y, y) has a unique solution in the ball B(0,1 ). In particular we have

kyk ≤ 2ky0k.

Proof. We start by defining the ball B(0, R) with

R = 1 −p1 − 4ηky0k

2η .

We will prove that

T (y) = y0+ B(y, y)

is a contraction in B(0, R). First of all let us prove that T : B(0, R) 7→ B(0, R),

kT (y)k ≤ ky0k + kB(y, y)k ≤ ky0k + ηkyk2 ≤ ky0k + ηR2 = R, where the last equality comes from the definition of R.

(19)

1.2. LITTLEWOOD-PALEY THEORY 7 Now we consider y1 and y2 and we compute

kT (y1) − T (y2)k = kB(y1, y1− y2) + B(y1− y2, y2)k

≤ η(ky1k + ky2k)ky1− y2k ≤ 2ηRky1− y2k, since 2ηR < 1 then we obtain that T is a contraction.

Thanks to Banach-Caccioppoli-Picard theorem we can state that the equation

y = y0+ B(y, y)

has a unique solution in the ball B(0, R). Moreover from this equation we get

kyk ≤ ky0k + ηkyk2 and so

kyk(1 − ηkyk) ≤ ky0k.

Now since kyk ≤ R ≤ 1

2η, we finally obtain that kyk ≤ 2ky0k.

To prove uniqueness in the ball B(0,1 ) let y = ye 0+ B(y,e y) be another solution,e with y ∈ B(0,e 1 ). We have

ky −eyk = kB(y, y −ey) + B(y −ey,ey)k

≤ η(kyk + keyk)ky −eyk ≤ η(R + 1

2η)ky −eyk < ky −eyk.



1.2 Littlewood-Paley theory

In this paragraph we briefly introduce Littlewood-Paley theory, showing some results concerning Sobolev and H¨older spaces

(20)

Let ϕ ∈ C0(R), 0 ≤ ϕ ≤ 1, ϕ0(ξ) = 1 if |ξ| ≤ 1110 and ϕ0(ξ) = 0 if |ξ| ≥ 1910.

For ν ∈ N, we consider ϕν(ξ) = ϕ(2−νξ) and we define byϕeν(x) the inverse Fourier transform of ϕν(ξ).

We define the following operators

S−1u = 0, Sνu =ϕeν ∗ u = ϕν(D)u, for ν ≥ 0,

0u = S0u, ∆νu = Sνu − Sν−1u, for ν ≥ 1.

Remark.

The Littlewood-Paley decomposition can be defined in a “homogeneous” way as follows

νu = Sνu − Sν−1u for ν ∈ Z, (1.12) this definitionis will be used in the definition of homogenous spaces.

We now report three proposition without proof concerning characterization of Sobolev and H¨older spaces through Littlewood-Paley decomposition:

Proposition 1 Let s ∈ R. A tempered distribution u belongs to Hs if and only if the following two conditions hold:

• for every ν ≥ 0, ∆νu ∈ L2

• the sequence δν = 2νsk∆νukL2 belongs to `2

Moreover there exists Cs > 1 such that, for every u ∈ Hs,

1

CskukHs ≤ X

ν

δν2

!12

≤ CskukHs. (1.13)

Proposition 2 Let s ∈ R and let R > 2. Let (uk)k a sequence of L2 functions such that

i) suppub0 is contained in {|ξ| ≤ R} and suppubkis contained in {R12k≤ |ξ| ≤ R2k} for k ≥ 1.

ii) the sequence (δk)k with δk = 2kskukkL2 is in `2.

(21)

1.2. LITTLEWOOD-PALEY THEORY 9 Then P

kuk is a converging series in Hs with sum u ∈ Hs and the norm in Hs of u is equivalent to the norm in `2 of (δk)k (that is there exists Cs such that (1.13) holds.

If s > 0, it is sufficient to suppose that supp buk is contained in {|ξ| ≤ R2k} for every k ≥ 1.

Proposition 3 A bounded function a belongs to C1, (with 0 <  < 1), the space of H¨older functions of indices 1, , if and only if the sequence (αk)k, with αk = 2k(1+)k∆kakL, is in `.

Moreover there exists C> 1, such that, for every a ∈ C1,, 1

CkakC1, ≤ sup

k

αk ≤ CkakC1,. (1.14)

Paraproduct

We now introduce Bony’s paraproduct, giving some results that can be obtained using this tool. Let’s start with a definition.

Definition. Let a ∈ L and u ∈ Hs. Bony’s paraproduct Tau is defined as follows

Tau =

+∞

X

ν=3

Sν−3a∆νu.

The following two propositions concern with some mapping properties of the para- product:

Proposition 4 i) Let s ∈ R and a ∈ L. Then Ta: Hs → Hs and kTaukHs ≤ CskakLkukHs.

ii) Let  ∈]0, 1[, a ∈ C1,, s ∈]0, 1 + ]. Then u 7→ au − Tau maps H−s in H1−s and kau − TaukH1−s ≤ CkakC1,kukH−s.

Proof. Since the result of i) is classical and can be found in [B] we give only the proof of ii). We have

(22)

au − Tau =

+∞

X

k=3

kaSk−3u

| {z }

(A)

+

+∞

X

k=3

( X

j ≥ 0

|j − k| ≤ 2

ka∆ju)

| {z }

(B)

,

regarding (A) we can observe that the support of the Fourier transform of ∆kaSk−3u is contained in {2k−2 ≤ |ξ| ≤ 2k+2} and one has

k∆kaSk−3ukL2 ≤ k∆kakLkSk−3ukL2

≤ CkakC1,2−k(1+)

k−3

X

j=0

2jsδj,

where (δj)j is in `2 and k(δj)jk`2 = CskukH−s. This yields

k∆kaSk−3ukL2 ≤ CkakC1,2−k(1+−s)

k−3

X

j=0

2−(k−j)sδj.

Now, defining eδk =

k−3

X

j=0

2−(k−j)sδj, as a consequence of Young inequality for convo- lution in `p, we have that (eδk)k∈ `2 and

k(eδk)kk`2 ≤ 2k(δk)kk`2. We easily obtain that (A) ∈ H1−s and

k(A)kH1−s ≤ CkakC1,kukH−s. Considering now (B) we have

(B) =

+∞

X

k=3

(∆ka∆k−2u + ∆ka∆k−1u + ∆ka∆ku + ∆ka∆k+1u + ∆ka∆k+2u).

(23)

1.2. LITTLEWOOD-PALEY THEORY 11

Every term can be treated analogously. We take the term

+∞

X

k=3

ka∆ku as an ex- ample. The support of its Fourier transform is contained in {|ξ| ≤ 2k+2} and

k∆ka∆kukL2 ≤ k∆kakLk∆kukL2 CkakC1,2−k(1+)2−ksδk, where as before (δj)j ∈ `2 and k(δj)jk`2 = CskukH−s.

As a consequence

+∞

X

k=3

ka∆ku ∈ H1−s and

k

+∞

X

k=3

ka∆kukH1−s ≤ CkakC1,kukH−s.

 Remark. In the last part, since the support of the Fourier transform of ∆ka∆ku is contained in a ball (and not in a ring) we can apply the second part of Proposition 2 and so it is necessary that 1 +  − s > 0. On the contrary it is easy to see that the result of Proposition 4 is valid also for s = 0.

The next result will be necessary in dealing with the backward parabolic operator of the thermoelastic system in Chapter 5.

Proposition 5 Let  ∈]0, 1[, s ∈]0, 1 + ], a ∈ C1,. Then

+∞

X

ν=0

2−2νsk∂x([∆ν, Ta]∂xu)k2L2

!12

≤ CkakC1,kukH1−s,

for every u ∈ H1−s.

([A, B] denotes the commutator of A and B, i.e. [A, B]u = A(B(u)) − B(A(u)).) Proof. We start by observing that

[∆ν, Ta]w =

+∞

X

K=3

[∆ν, Sk−3a]∆kw,

so

(24)

x([∆ν, Ta]∂xu) = ∂x

+∞

X

K=3

[∆ν, Sk−3a]∆k(∂xu)

! .

 From the support of the Fourier transform of [∆ν, Sk−3]∆kw we can obtain that [∆ν, Sk−3]∆kw is identically zero if |k − ν| ≥ 4.

This allows us to infer that the sum in k reduces at most to 7 terms, so

x([∆ν, Ta]∂xu) = ∂x([∆ν, Sν−6a]∆ν−3(∂xu)) + . . . + ∂x([∆ν, Sνa]∆ν+3(∂xu)).

The support of Fourier transform of each of these terms is contained in a ball proportional to 2ν. We consider one of these terms: one has

k∂x([∆ν, Sν−3a]∆ν(∂xu))kL2 ≤ C2νk[∆ν, Sν−3a]∂x(∆νu)kL2

≤ C2νkakLipk∆νukL2,

where the first inequality comes from Bernstein’s inequality, while the second de- rives from theorem 35 in Coifman and Meyer ([CM], see also [Ta]). The Proposition follows from this inequality and proposition 1.

1.3 Modulus of continuity

Let µ : [0, 1] → [0, 1] continuous, concave and strictly increasing function, with µ(0) = 0. We say thatµ is a modulus of continuity. Let I ⊆ R and f : I → R. We define f ∈ Cµ(I, R) if f ∈ L(I, R) and

sup

0 < |t − s| < 1 t, s ∈ I

|f (t) − f (s)|

µ(|t − s|) < +∞.

For example if µ(s) = s then Cµ= Lip.

In the following elementary proposition we collect some useful results:

Proposition 6 Let µ be a modulus of continuity. Then

(25)

1.3. MODULUS OF CONTINUITY 13

• µ(s) ≥ sµ(1) for all s ∈ [0, 1];

• the function s 7→ µ(s)

s is decreasing in ]0, 1];

• there exists lim

s→0+

µ(s) s ;

• the function σ 7→ µ(1σ)

(σ1) is increasing in [1, +∞];

• the function σ 7→ 1

σ2µ(σ1) is decreasing in [1, +∞].

We remark that if sup

s∈]0,1]

µ(s)

s < +∞, then there exists C > 0 such that µ(s) ≤ Cs for any s ∈ [0, 1]: so Cµ = Lip. Moreover, if Cµ6= Lip, then lim

s→0+

µ(s)

s = +∞.

We now define the so called Osgood condition.

Definition. (Osgood condition)

Let µ be a modulus of continuity. We say that µ satisfies the Osgood condition if:

Z 1 0

1

µ(s)ds = +∞.

From the modulus of continuity we will now define a weight function, that will play an important role in the study of the thermoelastic system.

Let µ be a modulus of continuity satisfying the Osgood condition, we define φ(t) =

Z 1

1 t

1 µ(s)ds,

φ ∈ C1 and it is strictly increasing. From the Osgood condition we obtain that φ([1, +∞[) = [0, +∞[ and φ0(t) = (t2µ(11

t)) > 0 for any t ∈ [1, +∞[. We define:

Φ(τ ) = Z τ

0

φ−1(s)ds,

(26)

so Φ0(τ ) = φ−1(τ ), then

τ →∞lim Φ0(τ ) = +∞.

Moreover

Φ00(τ ) = (Φ0(τ ))2µ

 1 Φ0(τ )



, (1.15)

for every τ ∈ [0, +∞[ and since σ 7→ σµ(σ1) is an increasing funtion on [1, +∞[, we obtain that:

τ →+∞lim Φ00(τ ) = lim

τ →+∞0(τ ))2µ

 1 Φ0(τ )



= +∞. (1.16)

To conclude this section we introduce a property of µ(s) that will be fundamental in the last chapter, where we will also give an example of a class of non - Lipschitz functions whose modulus of continuity satisfies this condition

Definition. ( ? condition)

Let µ be a modulus of continuity and let Φ defined as above: we say that µ satisfies the ? condition if there exists 0 < a < 1 such that

s→+∞lim

Φ0(s)

0(as))2 = 0. (?)

We remark that since Φ0 is an increasing function then, if µ satisfies this condition with a0, the same condition holds for every 0 < a0 < a < 1.

1.4 Functional spaces

In this section we introduce some functional spaces used in this thesis, focusing in particular on the less known ones.

(27)

1.4. FUNCTIONAL SPACES 15

Besov spaces

We start by introducing Besov spaces, using the definition of the operators ∆jgiven in section 1.2. We start with the inhomogeneous version: we say that f ∈ Bp,qs if

kf kBsp,q = X

j∈N

2qjsk∆jf kqLp

!1q

< +∞,

while for the homogeneous version we say that f ∈ ˙Bp,qs if

kf kB˙sp,q = X

j∈Z

2qjsk∆jf kqLp

!1q

< +∞,

where in this second case the operators ∆j are defined as in 1.12. We remark that in the case p = q = 2 we have that

B2,2s = Hs where Hs is the Sobolev space of index s.

In this work we will also use an anistropic version of Besov spaces, namely the space B0,s whose norm is defined as follows

kf kB0,s =X

j∈Z

2jsk∆vjf kL2 < +∞,

where the operators ∆vj define a monodimensional version of Littlewood-Paley decomposition.

The space fLp

In chapter 3 we will use the space fL(R+, B0,12), so it is worth to define fLp spaces and remark its connections with the space Lp, using the space B0,12 for example.

We say that u belongs to fLp(R+, B0,12) if

kuk

Lfp(R+,B0, 12) =X

j∈Z

2j2k∆vjukLp(R+,L2(R)) < +∞.

(28)

This definition slightly differs from the usual space Lp(R+, B0,12), that is defined as follows

kukLp(R+,B0, 12)=

X

j∈Z

22jk∆vjukL2(R)) Lp(R+)

< +∞,

and it is easy to prove that

kukLp(R+,B0, 12) ≤ kuk

Lfp(R+,B0, 12).

(29)

Chapter 2

Mild solutions for the

Navier-Stokes-Coriolis system:

the ˙ H 1 2 case

In this section we investigate the existence and uniqueness of mild solutions for the Navier-Stokes-Coriolis equations with Sobolev initial data. We consider the problem

tu − ν∆u + (u · ∇)u + Ωe3 × u + ∇p = 0

∇ · u = 0 u(0) = u0,

(2.1)

where u is the velocity field, p the pressure, ν the viscosity and Ω the Coriolis parameter.

Using the divergence free condition we can recast the system in the following way

tu − ν∆u + ∇ · (u ⊗ u) + Ωe3× u + ∇p = 0

∇ · u = 0 u(0) = u0,

(2.2)

(30)

2.1 Technical results

In this section we investigate some properties of the solution of the linear problem in the case of ˙Hs initial data.

Lemma 2 Let v0 ∈ ˙Hs and f ∈ L2([0, T ]; ˙Hs−1). We set v(t) = G(t)v0+

Z t 0

G(t − s)f (s)ds.

Then

v ∈

+∞

\

p=2

Lp([0, T ]; ˙Hs+2p)

!

∩ C([0, T ]; ˙Hs).

Moreover

(i) kv(t)k2H˙s + 2ν Z t

0

k∇v(t0)k2H˙sdt0 = kv0k2H˙s+ 2 Z t

0

< f (t0), v(t0) >s dt0;

(ii) Z

R3

|ξ|2s( sup

0≤t0≤T

|bv(t0)|)2dξ ≤ 9



kv0kH˙s + 1

√2νkf kL2([0,T ]; ˙Hs−1)

2

;

(iii) kv(t)k

Lp([0,T ];Hs+ 2p) ≤ 3 (ν)1p



kv0kH˙s + 1

√νkf kL2([0,T ]; ˙Hs−1)

 . Proof. One has

bv(t, ξ) =

 cos

 Ωtξ3

|ξ|



Id + sin

 Ωtξ3

|ξ|

 R(ξ)



e−ν|ξ|2tbv0(ξ)

+ Z t

0

 cos



Ω(t − s)ξ3

|ξ|



Id + sin



Ω(t − s)ξ3

|ξ|

 R(ξ)



e−ν|ξ|2(t−s)f (s, ξ)dsb

so we can write

|v(t, ξ)| ≤ 3|b bv0(ξ)| + Z t

0

3e−ν|ξ|2(t−s)| bf (t − s, ξ)|ds

≤ 3 |bv0(ξ)| +

Z t 0

e−2ν|ξ|2(t−s)ds

12 Z t 0

| bf (t − s, ξ)|2ds

12!

≤ 3 |bv0(ξ)| + 1

p2ν|ξ|2k bf (·, ξ)kL2([0,t])

! ,

(31)

2.1. TECHNICAL RESULTS 19 and then

|ξ|2s

 sup

0≤t0≤t

|bv(t0, ξ)|

2

≤ 9 |vb0(ξ)| + 1

p2ν|ξ|2k bf (·, ξ)kL2([0,t])

!2

|ξ|2s,

and finally we obtain (ii), in fact

Z

R3

|ξ|2s

 sup

0≤t0≤t

|bv(t0, ξ)|

2

!12

≤ 3k|bv0(ξ)| + 1

p2ν|ξ|2k bf (·, ξ)kL2([0,t])kH˙s

≤ 3kbv0kH˙s+ 3

√2ν

Z

R3

k bf (·, ξ)k2L2[0,t]|ξ|2s−2

12

≤ 3



kbv0kH˙s+ 1

√2νkf kL2([0,t], ˙Hs−1)

 .

Now we can deduce that v ∈ C([0, T ], ˙Hs). First we observe that form (ii) just obtained, one has v ∈ L([0, T ], ˙Hs) in fact

Z

R3

|ξ|2s|bv(t, ξ)|2dξ ≤ Z

R3

|ξ|2s

 sup

0≤t≤T

|bv(t, ξ)|2

 dξ

≤ 9



kbv0kH˙s+ 1

√2νkf kL2([0,t], ˙Hs−1)

2

,

the continuity is consequence of Lebesgue dominated convergence theorem, since

|ξ|2s|bv(t, ξ)|2 ≤ 18



|bv0(ξ)|2+ 1 2ν|ξ|2

Z T 0

| bf (s, ξ)|2ds



|ξ|2s.

Let us now prove (i). Suppose first that v ∈ L2([0, T ]; ˙Hs+1) ∩ C([0, T ], ˙Hs); since we have

( ∂tv − ν∆v + Ωe3× v = f − ∇p v(0) = v0,

then

tbv + ν|ξ|2bv + Ωe3×bv = bf − c∇p,

Riferimenti

Documenti correlati

We introduce a Gel’fand-Levitan-Marchenko theory, discuss some energy oscillation phenomena and provide some physical applications of an appropriate trace formula

Thus, the main objective of this research is to provide evidence on how a geographical information system (GIS) can be used to support health care management in the reorganization

The production of pancreatic beta cells producing insulin from stem cells in vitro may provide an unprecedented source of cells leading to the development of drugs

Finally, Section 4 is devoted to the proof of the Carleman estimate needed to deduce our uniqueness

More generally, the physical –evanescent mixing may first start at order L in perturbation theory, in which case the evanescent operators affect the anomalous dimension starting at

The loss of regularity for the coefficients at a single point is widely considered, e.g., in the case of well-posedness in the Cauchy problem for second-order hyperbolic operators of

The CFD model was applied in order to perform a parametric study involving dif- ferent aspects related to the machine: regenerator properties, working fluid, mean cycle

lento veloce debole forte piccolo grande stretto largo cattivo buono brutto bello chiuso aperto ricco povero basso alto corto lungo morto vivo femminile maschile vuoto pieno amaro