Chapter 2

Fluid mechanics & Orlicz spaces

Abstract

We extend some classical tools from fluid mechanics – Korn's inequality, the Bogovskiĭ  operator and the pressure recovery – to the setting of Orlicz spaces. As a special case the known LpImage-theory is included as well as the case of Orlicz spaces generated by a nice Young function (i.e., under Δ2Image and 2Image condition). In the general case there is some loss of integrability, for instance in the limit cases LlogLL1Image and LExp(L)Image. The results are shown to be optimal in the sense of Orlicz spaces.

Keywords

Orlicz spaces; Divergence equation; Bogovskiĭ operator; Negative norm theorem; Pressure recovery; Korn's inequality

A crucial tool in the mathematical approach to the behaviour of Newtonian fluids is Korn's inequality: given a bounded open domain GRdImage, d2Image, with Lipschitz boundary ∂G we have

G|v|2dx2G|ε(v)|2dx

Image (2.0.1)

for all vW1,20(G)Image. For smooth functions with compact support (2.0.1) can be shown by integration by parts. The general case is treated by approximation. A first proof was given by Korn in [104]. We note that variants of Korn's inequality in L2Image have been established by Courant and Hilbert [53], Friedrichs [84], Èidus [70] and Mihlin [114]. Many problems in the mathematical theory of generalized Newtonian fluids and mechanics of solids lead to the following question (compare for example the monographs of Málek, Nečas, Rokyta and Růžička [111], of Duvaut and Lions [66] and of Zeidler [143]): is it possible to bound a suitable energy depending on vImage by the corresponding functional of ε(v)Image, that is

G|v|pdxc(p,G)G|ε(v)|pdx

Image (2.0.2)

for functions vW1,p0(G)Image? As shown by Gobert [91,92], Nečas [119], Mosolov and Mjasnikov [116], Temam [138] and later by Fuchs [74] this is true for all 1<p<Image (we remark that the inequality fails in the case p=1Image, see [120] and [52]).

A first step in the generalization of (2.0.2) is mentioned in [4]: Acerbi and Mingione prove a variant for the Young function

A(t)=(1+t2)p22t2.

Image

More precisely, they show that

vLA(G)c(φ,G)ε(v)LA(G)

Image (2.0.3)

for all functions vW1,A0(G)Image. Although they only consider a special case they provide tools for much more general situations. Note that they only obtain inequalities in the Luxembourg-norm which is not appropriate in many situations (for example in regularity theory, see [35]). A general theorem is proved in [64], namely that

GA(|v(v)G|)dxc(A,G)GA(|ε(v)(ε(v))G|)dx

Image (2.0.4)

for all vW1,A(G)Image, where A is a Young function satisfying the Δ2Image- and 2Image-condition. Furthermore, Fuchs [75] obtains (2.0.4) for functions with zero traces and the same class of Young functions by a different approach. It is shown in [32] that the Δ2Image- and 2Image-condition are also necessary for the inequality (2.0.4). We remark that the constitutive law

S=A(|ε(v)|)|ε(v)|ε(v)

Image

for a Young function A is a quite general model to describe the motion of generalized Newtonian fluids (see, i.e., [35], [25] and [59]).

In order to characterize the behaviour of Prandtl–Eyring fluids (see Chapter 4) Eyring [69] suggested the constitutive law

S=DW(ε(u)),W(ε)=h(|ε|)=|ε|log(1+|ε|).

Image (2.0.5)

This leads in a natural way to the question about Korn's inequality in the space Lh(G)Image. Since we have

˜h(t)t(exp(t)1),

Image

the 2Image-condition fails in this case, hence the results mentioned above do not apply. So the following question remains: given some integrability of the symmetric gradient – in the sense of Orlicz spaces – what is the best integrability for the full gradient we can expect?

A second fundamental question in fluid mechanics is the recovery of the pressure. It is common (and very useful) to study pressure-free formulations of (generalized) Navier–Stokes equations. So one starts by finding a velocity field which solves the corresponding system in the sense of distributions on divergence-free test-functions. Afterwards there is the question about the existence of the pressure function in order to have a weak solution in the sense of distributions.

Let us be a bit more precise and consider the equation

GH:φdx=0for allφC0,div(G),

Image (2.0.6)

where H is an integrable function (in case of a stationary generalized Navier–Stokes equation we have H=S(ε(v))+Δ1fvvImage). Secondly, the pressure π is reconstructed in the sense that

GH:φdx=Ωπdivφdxfor allφC0(G).

Image (2.0.7)

The existence of a pressure in the sense of distributions is a consequence of the classical theorem by De Rahm (see [131] for an appropriate version). It is also well-known that – if 1<p<Image – then HLp(G)Image implies πLp(G)Image. This result breaks down in the limit cases. Again motivated by the Prandtl–Eyring model the following question remains: given a function H solving (2.0.6) – located in some Orlicz space – what is the optimal integrability of the pressure in (2.0.7)?

In classical LpImage-spaces, both questions raised above can be answered by Nečas' negative norm theorem [119]. The negative Sobolev norm of the distributional gradient of a function uL1(G)Image can be defined as

uW1,p(G)=supφC0(G)GudivφdxφLp(G)dx,

Image (2.0.8)

where 1pImage. In (2.0.8), and in similar occurrences throughout this chapter, we tacitly assume that the supremum is extended over all functions v which do not vanish identically. We remark that the quantity on the right-hand side of (2.0.8) agrees with the norm of ∇u, when regarded as an element of the dual of W1,p0(G)Image. Nečas showed that, if G is regular enough – a bounded Lipschitz domain, say – and 1<p<Image, then the Lp(G)Image norm of a function is equivalent to the W1,p(G)Image norm of its gradient. Namely, there exist positive constants C1=C1(G,p)Image and C2=C2(d)Image, such that

C1uuGLp(G)uW1,p(G)C2uuGLp(G)

Image (2.0.9)

for every uL1(G)Image.

Using the formula

Δu=divV(u),Vij(u)=2εD(u)(121d)(divu)I,

Image

where εD=ε1dtrεIImage, the proof of Korn's inequality based on (2.0.9) is elementary. Moreover, a combination of De Rahm's Theorem and (2.0.9) shows that if HLp(G)Image satisfies (2.0.6) then there is πLp(G)Image such that (2.0.7) holds.

In order to understand how Korn's inequality and the pressure recovery work in Orlicz spaces, we have to understand Orlicz versions of (2.0.9). Let A be a Young function, and let G be a bounded domain in RdImage. We define the negative Orlicz–Sobolev norm associated with A of the distributional gradient of a function uL1(G)Image as

uW1,A(G)=supφC0(G,Rn)GudivφdxφL˜A(G).

Image (2.0.10)

The alternative notation W1LA(G)Image will also occasionally be employed to denote the negative Orlicz–Sobolev norm W1,A(G)Image associated with the Orlicz space LA(G)Image. As (2.0.9) is known to break down in the limit cases, an Orlicz-version with the same Young function on both sides cannot hold in general. In fact, our Orlicz–Sobolev space version of the negative norm theorem involves pairs of Young functions A and B which obey the following balance conditions:

tt0B(s)s2dsA(ct)fort0,

Image (2.0.11)

and

tt0˜A(s)s2ds˜B(ct)fort0,

Image (2.0.12)

for some positive constant c. Note that the same conditions come into play in the study of singular integral operators in Orlicz spaces [47].

If either (2.0.11) or (2.0.12) holds, then A dominates B globally [48, Proposition 3.5]. In a sense, the assumptions (2.0.11) and (2.0.12) provide us with a quantitative information about how much weaker the norm LB(G)Image is than LA(G)Image. Under these assumptions a version of the negative norm theorem can be restored in Orlicz–Sobolev spaces.

Theorem 2.0.5

Let A and B be Young functions, fulfilling (2.0.11) and (2.0.12). Assume that G is a bounded domain with the cone property in RdImage, d2Image. Then there exist constants C1=C1(G,c)Image and C2=C2(d)Image such that

C1uuGLB(G)uW1,A(G)C2uuGLA(G)

Image (2.0.13)

for every uL1(G)Image. Here, c denotes the constant appearing in (2.0.11) and (2.0.12).

Remark 2.0.6

Inequality (2.0.13) continues to hold even if conditions (2.0.11) and (2.0.12) are just fulfilled for tt0Image for some t0>0Image, but with constants C1Image and C2Image depending also on A, B, t0Image and |G|Image. Indeed, the Young functions A and B can be replaced, if necessary, with Young functions which are equivalent near infinity and fulfil (2.0.11) and (2.0.12) for every t>0Image. Due to (1.2.11), such replacement leaves the quantities LA(G)Image, LB(G)Image and W1,A(G)Image unchanged, up to multiplicative constants depending on A, B, t0Image and |G|Image.

The situations when (2.0.11), or (2.0.12), holds with B=AImage can be precisely characterized. Membership of A to Δ2Image is a necessary and sufficient condition for (2.0.12) to hold with B=AImage [103, Theorem 1.2.1]. Therefore, under this condition, assumption (2.0.12) can be dropped in Theorem 2.0.5. On the other hand, A2Image if and only if ˜AΔ2Image. Hence membership of A to 2Image is a necessary and sufficient condition for (2.0.11) to hold with B=AImage. Thus, under this condition, assumption (2.0.11) can be dropped in Theorem 2.0.5. Particularly, if AΔ22Image, then both conditions (2.0.11) and (2.0.12) are fulfilled with B=AImage. Hence, we have the following corollary which also follows from the results of [64].

Corollary 2.0.1

Assume that G is a bounded domain with the cone property in RdImage, d2Image. Let A be a Young function in Δ22Image. Then there are two constants C=C(G,A)Image and C2=C2(d)Image such that

C1uuGLA(G)uW1,A(G)C2uuGLA(G)

Image (2.0.14)

for every uL1(G)Image.

A typical situation where condition (2.0.11) does not hold with B=AImage is when A grows linearly, or “almost linearly”, near infinity. In this case, A2Image. In fact, as already mentioned, the standard negative norm theorem expressed by (2.0.9) breaks down in the borderline case p=1Image. On the other hand, condition (2.0.12) fails, with B=AImage, if, for example, A has a very fast – faster than any power – growth. In this case, AΔ2Image. Loosely speaking, the norm LA(G)Image is now “close” to L(G)Image, and, as a matter of fact, equation (2.0.9) is not true with p=Image.

These, however, are not the only situations when (2.0.11), or (2.0.12), fail with B=AImage. For instance, there are functions A which neither satisfy the Δ2Image condition, nor the 2Image condition. Therefore, neither (2.0.12) nor (2.0.11) can hold with B=AImage. In those cases A(t)Image “oscillates” between two different powers tpImage and tqImage, with 1<p<q<Image. Functions of this kind are referred to as (p,q)Image-growth in the literature. Partial differential equations, and associated variational problems, whose nonlinearity is governed by this growth, have been extensively studied. In the framework of non-Newtonian fluids, they have been analysed in [22].

All the circumstances described above can be handled via Theorem 2.0.5. A few examples involving customary families of Young functions are presented hereafter.

Example 1

Assume that A(t)Image is a Young function equivalent to tplogα(1+t)Image near infinity, where either p>1Image and αRImage, or p=1Image and α1Image. Hence, if |G|<Image, then

LA(G)=LplogαL(G).

Image

Assume that G is a bounded domain with the cone property in RdImage. If p>1Image, then AΔ22Image, and hence Corollary 2.0.1 tells us that

C1uuGLplogαL(G)uW1LplogαL(G)C2uuGLplogαL(G)

Image (2.0.15)

for every uL1(G)Image. However, if p=1Image, then AΔ2Image, but A2Image. An application of Theorem 2.0.5 now implies

C1uuGLlogα1L(G)uW1LlogαL(G)C2uuGLlogαL(G)

Image (2.0.16)

for every uL1(G)Image. In particular,

C1uuGL1(G)uW1LlogL(G)C2uuGLlogL(G)

Image (2.0.17)

for every uL1(G)Image.

Example 2

Let β>0Image, and let A(t)Image be a Young function equivalent to exp(tβ)Image near infinity. Then

LA(G)=expLβ(G)

Image

if |G|<Image. One has that A2Image, but AΔ2Image. Theorem 2.0.5 ensures that, if G is a bounded domain with the cone property in RdImage, then

C1uuGexpLββ+1(G)uW1expLβ(G)C2uuGexpLβ(G)

Image (2.0.18)

for every uL1(G)Image. Moreover,

C1uuGexpL(G)uW1L(G)C2uuGL(G)

Image (2.0.19)

for every uL1(G)Image.

Our approach is based on a study of the Bogovskiĭ  operator [24] in Orlicz spaces in Theorem 2.1.7 which is already interesting itself. The Bogovskiĭ  operator is a solution operator to the divergence equation with respect to zero boundary conditions. The continuity of the Bogovskiĭ  operator implies the negative norm Theorem from which we can deduce both, the pressure recovery and Korn's inequality.

In Theorem 2.2.10 we give the precise statement of the pressure recovery in Orlicz spaces. In fact, HLA(G)Image implies πLB(G)Image where A and B are linked through (2.0.11) and (2.0.12). Moreover, the following inequality holds

GB(|π|)dxGA(C|HHG|)dx.

Image

Theorem 2.3.12 contains a version of Korn's inequality in general Orlicz spaces which says

GB(|u(u)G|)dxGA(C|ε(u)(ε(u))G|)dx.

Image

The final question which remains is the sharpness of the mentioned results. In Section 2.3 we are going to show that the balance conditions (2.0.11) and (2.0.12) are also necessary for a Korn's inequality. This implies that also the results about negative norms in Theorem 2.0.5 and the Bogovskiĭ  operator in Theorem 2.1.6 are optimal.

2.1 Bogovskiĭ operator

Our proof of Theorem 2.0.5 relies upon an analysis of the divergence equation

{divu=finG,u=0onG,

Image (2.1.20)

in Orlicz spaces which we analyse in the following. Subsequently, we set

C0,(G)={uC0(G):uG=0},LA(G)={uLA(G):uG=0},

Image

where uG=GudxImage denotes the mean value of the function u.

Theorem 2.1.6

Assume that G is a bounded domain with the cone property in RdImage, d2Image. Let A and B be Young functions fulfilling (2.0.11 and (2.0.12). Then there exists a bounded linear operator

BogG:LA(G)W1,B0(G)

Image (2.1.21)

such that

BogG:C0,(G)C0(G)

Image (2.1.22)

and

div(BogGf)=fin G

Image (2.1.23)

for every fLA(G)Image. In particular, there exists a constant C=C(G,c)Image such that

(BogGf)LB(G)CfLA(G)

Image (2.1.24)

and

GB(|(BogGf)|)dxGA(C|f|)dx

Image (2.1.25)

for every fLA(G)Image. Here, c denotes the constant appearing in (2.0.11) and (2.0.12).

Although it will not be used for our main purposes, we state in Theorem 2.1.7 below a result parallel to Theorem 2.1.6, dealing with a version of problem (2.1.20) in the case when the right-hand side of the equation is in divergence form. Namely,

{divu=divginG,u=0onG,

Image (2.1.26)

where g:GRdImage is a given function satisfying the compatibility condition (in a weak sense) that its normal component on ∂G vanishes. As a precise formulation of this condition we consider the space HA(G)Image of those vector-valued functions u:GRnImage for which the norm

uHA(G)=uLA(G,Rn)+divuLA(G)

Image (2.1.27)

is finite. We denote by HA0(G)Image its subspace of those functions uHA(G)Image whose normal component on ∂G vanishes, in the sense that

Gφdivudx=Guφdx

Image (2.1.28)

for every φC(G)Image. It is easy to see that both HA(G)Image and HA0(G)Image are Banach spaces.

Theorem 2.1.7

Assume that G is a bounded Lipschitz domain in RdImage, d2Image. Let A and B be Young functions fulfilling (2.0.11) and (2.0.12). Then there exists a bounded linear operator

EG:HA0(G)W1,B0(G)

Image (2.1.29)

such that

div(EGg)=divginG

Image (2.1.30)

for every gHA0(G)Image. In particular, there exists a constant C=C(G,c)Image such that

(EGg)LB(G)CdivgLA(G)

Image (2.1.31)

and

EGgLB(G)CgLA(G)

Image (2.1.32)

for every gHA0(G)Image. Here, c denotes the constant appearing in (2.0.11) and (2.0.12).

The proofs of Theorems 2.1.6 and 2.1.7 make use of a rearrangement estimate, which extends those of [18, Theorem 16.12] and [15], for a class of singular integral operators of the form

Tf(x)=limε0+{y:|yx|>ε}K(x,y)f(y)dyforxRd,

Image (2.1.33)

for an integrable function f:RdRImage. Here K(x,y)=N(x,xy)Image, where the kernel N:Rd×RdRImage fulfills the following properties:

N(x,λz)=λdN(x,z)forx,zRd;

Image (2.1.34)

Sd1N(x,z)dHd1(z)=0forxRd;

Image (2.1.35)

For every σ[1,)Image, there exists a constant C1Image such that

(Sd1|N(x,z)|σdHd1(y))1σC1(1+|x|)dforxRd,

Image (2.1.36)

where Ss1Image denotes the unit sphere, centered at 0 in RdImage, and Hs1Image stands for the (s1)Image-dimensional Hausdorff measure.

There exists a constant C2Image such that

|K(x,y)|C2(1+|x|)d|xy|sforx,yRd,xy,

Image (2.1.37)

and, if 2|xz|<|xy|Image, then

|K(x,y)K(z,y)|C2(1+|y|)d|xz||xy|d+1,

Image (2.1.38)

|K(y,x)K(y,z)|C2(1+|y|)d|xz||xy|d+1.

Image (2.1.39)

Theorem 2.1.8

Let G be a bounded open set in RdImage, and let K be a kernel satisfying (2.1.34)(2.1.39). If fL1(Rd)Image and f=0Image in RdGImage, then the singular integral operator T given by (2.1.33) is well defined for a.e. xRdImage, and there exists a constant C=C(C1,C2,d,diam(G))Image for which

(Tf)(s)C(1ss0f(r)dr+|G|sf(r)drr)fors(0,|G|).

Image (2.1.40)

As a consequence of Theorem 2.1.8, the boundedness of singular integral operators given by (2.1.33) between Orlicz spaces associated with Young functions A and B fulfilling (2.0.11) and (2.0.12) can be established.

Theorem 2.1.9

Let G, K and T be as in Theorem 2.1.8. Assume that A and B are Young functions satisfying (2.0.11) and (2.0.12). Then there exists a constant C=C(C1,C2,d,diam(G),c)Image such that

TfLB(G)CfLA(G),

Image (2.1.41)

and

GB(|Tf|)dxGA(C|f|)dx

Image (2.1.42)

for every fLA(G)Image. Here, c denotes the constant appearing in (7.0.1) and (5.3.12).

Proof

According to Lemma 1.2.1, if A and B are Young functions satisfying (2.0.11), then there exists a constant C=C(c)Image such that

1ss0φ(r)drLB(0,)CφLA(0,)

Image (2.1.43)

for every φLA(0,)Image. Moreover, if A and B fulfil (2.0.12), then there exists a constant C=C(c)Image such that

sφ(r)drrLB(0,)CφLA(0,)

Image (2.1.44)

for every φLA(0,)Image. Combining (2.1.40), (2.1.43) and (2.1.44), and making use of property (1.2.12) implies inequality (2.1.41).

As far as (2.1.42) is concerned, observe that, inequalities (2.0.11) and (2.0.12) continue to hold, with the same constant c, if A and B are replaced with kA and kB, where k is any positive constant. Thus, inequality (2.1.41) continues to hold, with the same constant C, after this replacement, whatever k is, namely

TfLkB(G)CfLkA(G)

Image (2.1.45)

for every fLA(G)Image. Now, given any such f, choose k=1GA(|f|)dxImage. The very definition of Luxemburg norm tells us that fLkA(G)1Image. Hence, by (2.1.45), TfLkB(G)CImage. The definition of Luxemburg norm again implies that GkB(|Tf|C)dx1Image, namely (2.1.42).  □

Proof of Theorem 2.1.8

Let R>0Image be such that GBR(0)Image, the ball centered at 0, with radius R. Fix a smooth function η:[0,)[0,)Image for which η=1Image in [0,3R]Image and η=0Image in [4R,)Image. Define

ˆN(x,z)=η(|x|)N(x,z)forx,zRd,ˆK(x,y)=ˆN(x,xy)forx,yRd.

Image

By properties (2.1.34)(2.1.39), one has that:

ˆN(x,λy)=λdˆN(x,z)forx,zRd;

Image (2.1.46)

Sd1ˆN(x,z)dHd1(z)=0forxRd;

Image (2.1.47)

for every σ[1,)Image, there exists a constant ˆC1=ˆC1(C1,σ,R,d)Image such that

(Sd1|ˆN(x,z)|σdHd1(z))1σˆC1forxRd,

Image (2.1.48)

where C1Image is the constant appearing in (2.1.36); there exists a constant ˆC2=ˆC2(C2,R,d)Image for which

|ˆK(x,y)|ˆC2|xy|dforx,yRd,xy,

Image (2.1.49)

and, if xRdImage, yGImage and 2|xz|<|xy|Image, then

|ˆK(x,y)ˆK(z,y)|ˆC2|xz||xy|d+1,

Image (2.1.50)

|ˆK(y,x)ˆK(y,z)|ˆC2|xz||xy|d+1,

Image (2.1.51)

where C2Image is the constant appearing in (2.1.37)(2.1.39). Define

ˆTεf(x)={y:|yx|>ε}ˆK(x,y)f(y)dy,ˆTSf(x)=supε>0|ˆTε(f)(x)|.

Image

Inequality (2.1.40) will follow if we prove that

(ˆTSf)(s)C(1ss0f(r)dr+|G|sf(r)drr)fors(0,)

Image (2.1.52)

for some constant C=C(C1,C2,d,R)Image, and for every fL1(Rd)Image for which f=0Image in RdBR(0)Image. A proof of inequality (2.1.52) can be accomplished along the same lines as that of Theorem 1 of [15], which in turn relies upon similar techniques as in [51]. For completeness, we give the details of the proof hereafter.

The key step in the derivation of (2.1.52) consists in showing that, for every γ(0,1)Image, there exists a constant C=C(C1,C2,γ,d,R)Image such that

(ˆTSf)(s)C(Mf)(γs)+(ˆTSf)(2s)fors(0,)

Image (2.1.53)

for every fL1(Rd)Image with f=0Image in RdBR(0)Image. Fix s>0Image, and define

E={xRd:ˆTSf(x)>(ˆTSf)(2s)}.

Image

Then, there exists an open set UEImage for which |U|3sImage. By Whitney's covering theorem, there exist a family of disjoint cubes {Qk}Image such that U=k=1QkImage, k=1|Qk|=|U|3sImage, and

diam(Qk)dist(Qk,RdU)4diam(Qk)forkN.

Image

The operator ˆTSImage is of weak type (1,1)Image, namely, there exists a constant CImage such that

|{xRd:ˆTSf(x)>λ}|CλfL1(Rd)

Image (2.1.54)

for fL1(Rd)Image. The proof of (2.1.54) follows from classical arguments: By (2.1.51) we have for all r>0Image, yRdImage and all xBr(z)Image that

|yx|2r|ˆK(y,x)ˆK(y,z)|dyC.

Image

So ˆKImage satisfies condition (10) in [134, p. 33]. By [134, Cor. 1, p. 33] we obtain

|{xRd:ˆTεf(x)>λ}|CλfL1(Rd),

Image (2.1.55)

where C does not depend on ε. Now (2.1.55) implies (2.1.54) by taking the supremum in ε.

We now show that there exists a constant CImage such that

|{xQk:ˆTSf(x)>CMf(x)+(ˆTSf)(2s)}|1γ3|Qk|forkN.

Image (2.1.56)

Fix any kNImage, choose xkRdUImage such that dist(xk,Qk)4diam(Qk)Image, and denote by Q the cube, centered at xkImage, with diam(Q)=20diam(Qk)Image. Define

g=fχQ,h=fχRdQ,

Image

so that f=g+hImage. If we prove that there exist constants C1Image and C2Image such that

ˆTSh(x)C1Mf(x)+(ˆTSf)(2s)forxQk,

Image (2.1.57)

and

|{xQk:ˆTSg(x)>C2Mf(x)}|1γ3|Qk|,

Image (2.1.58)

then (2.1.56) follows with C=C1+C2Image. Consider (2.1.58) first. Let C2Image be a constant for which C|Q|C21γ3|Qk|Image. Let λ=C2|Q|Q|g|dxImage. Since C2Mf(x)λImage for xQkImage, an application of (2.1.54) with this choice of λ tells us that

|{xQk:ˆTSg(x)>C2Mf(x)}||{ˆTSg(x)>λ}|CλQ|g|dxC|Q|C21γ3|Qk|,

Image

namely (2.1.58). In order to establish (2.1.57), it suffices to prove that, for every ε>0Image,

|ˆTεh(x)|C1Mf(x)+ˆTSf(xk)forxQk.

Image (2.1.59)

Indeed, since xkUImage, we have that ˆTSf(xk)(ˆTSf)(2s)Image, and hence (2.1.59) implies (2.1.57). We may thus focus on (2.1.59). Fix ε>0Image, and set r=max{ε,dist(xk,RdQ)}Image. Observe that r>10diam(Qk)Image. Given any xQkImage, define V=Bε(x)Bε(xk)Image. One has that

|ˆTεh(x)|=|{y:|yx|>ε}ˆK(x,y)h(y)dy||{y:|yxk|>ε}ˆK(x,y)h(y)dy|+V|ˆK(x,y)h(y)|dy.

Image (2.1.60)

Observe that, if ysupphImage, then |xy|>r2Image and hence 1|xy|d<2drdImage. Thus, due to (2.1.49),

|ˆK(x,y)|ˆC2rd.

Image

Moreover, VB3r(x)Image. Therefore, there exists a constant ˆCImage such that

V|ˆK(x,y)h(y)|dyˆCB3r(x)|h(y)|dyˆCMh(x)ˆCMf(x).

Image (2.1.61)

On the other hand,

|{y:|yxk|>ε}ˆK(x,y)h(y)dy||{y:|yxk|>r}ˆK(x,y)h(y)dy||{y:|yxk|>r}ˆK(xk,y)f(y)dy|+{y:|yxk|>r}|ˆK(xk,y)ˆK(x,y)||f(y)|dyˆTS(xk)+{y:|yxk|>r}|ˆK(xk,y)ˆK(x,y)||f(y)|dy,

Image (2.1.62)

where the first inequality holds since h(y)=0Image in {y:|yxk|r}Image if r=dist(xk,RdQ)Image, and trivially holds (with equality) if r=εImage. Since 2|xxk||xy|Image in the last integral in (2.1.62), and f vanishes in RdBR(0)Image, by (2.1.50)

|ˆK(xk,y)ˆK(x,y)|ˆC2|xkx||xy|d+1ˆC2diam(Qk)|xy|d+1.

Image

Hence,

{y:|yxk|>r}|ˆK(xk,y)ˆK(x,y)||f(y)|dy{y:|yx|>diam(Qk)}|f(y)|diam(Qk)|xy|d+1dy˜CMf(x)

Image (2.1.63)

for some constant ˜CImage. Note that, in the first inequality, we made use of the inclusion {y:|yxk|>r}{y:|yx|>diam(Qk)}Image, which holds since |xxk|<5diam(Qk)Image, and 10diam(Qk)<rImage.

Combining inequalities (2.1.60)(2.1.63) implies (2.1.59). Inequality (2.1.56) is fully established. Via summation in kQkImage, we obtain from (2.1.56) that

|{xRd:ˆTSf(x)>ˆCMf(x)+(ˆTSf)(2s)}|(1γ)s.

Image (2.1.64)

Combining (2.1.64) with the inequality

|{xRd:Mf(x)>(Mf)(γs)}|γs

Image (2.1.65)

tells us that

|{xRd:ˆTSf(x)>ˆC(Mf)(γs)+(ˆTSf)(2s)}||{ˆTSf(x)>ˆCMf(x)+(ˆTSf)(2s)}|+|{Mf(x)>(Mf)(γs)}|s.

Image

Hence (2.1.53) follows, by the very definition of decreasing rearrangement.

Starting from inequality (2.1.53) we apply the iteration argument from [15, Lemma 3.2] and obtain for γ=1/2Image

(ˆTSf)(s)Ck=0(Mf)(2k1s)+lims(ˆTSf)(s).

Image

Therefore, we have for all f satisfying lims(ˆTSf)(s)=0Image that

(ˆTSf)(s)Ck=2(Mf)(2k1s)+2C(Mf)(s2).

Image

Since

(Mf)(2k1s){2k2sσ2k1s}(Mf)(σ)dσσ

Image

we conclude

(ˆTSf)(s)Cs(Mf)(σ)dσσ+2C(Mf)(s2).

Image (2.1.66)

Now we fix s>0Image and assume that s(Mf)(σ)dσσImage is finite. Since each f has compact support, ˆTSf(x)Image converges to zero for all k as |x|Image (recall (2.1.49) and the definition of ˆTSImage). Therefore, we have that lims(ˆTSf)(s)=0Image for every k. Hence (2.1.66) implies

(ˆTSf)(s)Cs(Mf)(σ)dσσ+2C(Mf)(s2)C(sf(r)drr+f(s2))C(sf(r)drr+2sss2f(r)dr)C(sf(r)drr+1ss0f(r)dr)

Image

which shows (2.1.40).  □

Lemma 2.1.1

Let G be a bounded domain with the cone property in RdImage, with n2Image. Then there exist NNImage and a finite family {Gi}i=0,NImage of domains which are starshaped with respect to balls, for which G=Ni=0GiImage. Moreover, given fLA(G)Image, there exist fiLA(G)Image, i=0,NImage, such that fi=0Image in GGiImage,

f=Ni=0fi

Image

and

fiLA(G)CfLA(G)fori=0,,N,

Image (2.1.67)

for some constant C=C(G)Image.

Proof, sketched

Any bounded open set with the cone property can be decomposed into a finite union of Lipschitz domains [5, Lemma 4.22]. On the other hand, any Lipschitz domain can be decomposed into a finite union of open sets which are starshaped with respect to balls [85, Lemma 3.4, Chapter 3]. This proves the existence of the domains {Gi}i=0,NImage as in the statement. The same argument as in the proof of [85, Lemma 3.2, Chapter 3] then enables us to construct the desired family of functions fiImage on G, i=1,,NImage, according to the following iteration scheme. We set Di=Nj=i+1GjImage, g0=fImage, and, for i=1,,N1Image,

gi(x)={(1χGiDi(x))gi1(x)χGiDi(x)|GiDi|DiGigi1(y)dyif xDi,0otherwise,

Image (2.1.68)

and

fi(x)={gi1(x)χGiDi(x)|GiDi|Gigi1(y)dyif xGi,0otherwise.

Image (2.1.69)

Observe that, since G is connected, we can always relabel the sets GiDiImage in such a way that |GiDi|>0Image for i=1,,N1Image. Finally, we define

fN=gN1.

Image (2.1.70)

The family {fi}Image satisfies the required properties. The only nontrivial property is (2.1.67). To verify it, fix i, and observe that, by (2.1.69), the second inequality in (1.2.10), inequality (1.2.3), and inequality (1.2.7)

fiLA(G)gi1LA(G)(1+2|GiDi|1LA(GiDi)1L˜A(Gi))=gi1LA(G)(1+2|GiDi|A1(1/|GiDi|)1˜A1(1/|Gi|))gi1LA(G)(1+4˜A1(1/|GiDi|)˜A1(1/|Gi|))gi1LA(G)(1+4|Gi||GiDi|).

Image (2.1.71)

On the other hand, by (2.1.68) and a chain similar to (2.1.71), one has that

gi1LA(G)gi2LA(G)(1+2|Gi1Di1|1LA(Gi1Di1)1L˜A(Di1))gi2LA(G)(1+4˜A1(|1/Gi1Di1|)˜A1(|1/Di1|))gi2LA(G)(1+4|Di1||Gi1Di1|).

Image (2.1.72)

From (2.1.71), and an iteration of (2.1.72), one infers that

fiLA(G)(1+4|Gi||GiDi|)i1j=1(1+4|Dj||GjDj|)fLA(G),

Image

and (2.1.67) follows.  □

Proof of Theorem 2.1.6

By Lemma 2.1.1, it suffices to prove the statement in the case when G is a domain starshaped with respect to a ball BImage, which, without loss of generality, can be assumed to be centered at the origin and with radius 1. In this case, we are going to show that the (gradient of the) Bogovskiĭ  operator BogGImage, defined at a function fLA(G)Image is

BogGf(x)=Gf(y)(xy|xy|d|xy|ω(y+ζxy|xy|)ζd1dζ)dy

Image (2.1.73)

for xGImage. Here ω is any (nonnegative) function in C0(B)Image with Bωdx=1Image, agrees with a singular integral operator, whose kernel fulfills (2.1.34)(2.1.39), plus two operators enjoying stronger boundedness properties. If fC0(G)Image it is easy to see that the same is true for BogGfImage using the representations

BogGf(x)=Gf(y)(xy)(xy|xy|d1ω(y+ζ(xy))ζd1dζ)dy=Gf(y)(xy)(xy|xy|d0ω(y+ζ(xy)|xy|)(|xy|ζ)d1dζ)dy.

Image

Setting u=BogGfImage, we have

u(x)=Gf(y)N(x,y)dyforxG,

Image

where

N(x,y)=xy|xy|d|xy|ω(y+ζxy|xy|)ζd1dζforx,yG.

Image

By standard arguments we obtain (see [85, Proof of Lemma III.3.1] for more details)

jui(x)=Gf(y)jNi(x,y)dy+f(x)G(xy)i(xy)j|xy|2ω(y)dy.

Image

Computing jNImage we see that

jN=Kij(x,y)+Gij(x,y)

Image

with

Kij(x,y)=δij|xy|d0ω(x+ζxy|xy|)ζd1dζ+xiyi|xy|d+10jω(x+ζxy|xy|)ζddζ,

Image (2.1.74)

Gij(x,y)=xiyi|xy|d+10jω(x+ζxy|xy|)d1k=0(dk)|xy|kζdkdζ.

Image (2.1.75)

Now we want to justify the formula for a non-smooth function f. We claim that uW1,10(G)Image, and

uixj=Hijffor a.e.xG,

Image (2.1.76)

where HijImage is the linear operator defined at f as

(Hijf)(x)=GKij(x,y)f(y)dy+GGij(x,y)f(y)dy+f(x)G(xy)i(xy)j|xy|2ω(y)dyforxG,

Image (2.1.77)

for i,j=1,dImage. Here, KijImage is the kernel of a singular integral operator satisfying the same assumptions as the kernel K in Theorem 2.1.8. Moreover, the following holds

|Gij(x,y)|c|xy|d1forx,yRd,xy.

Image (2.1.78)

Computing the divergence based on (2.1.76) and (2.1.77) we see that

divu=Gdf(y)1ω(y+r(xy))rd1drdy+Gf(y)di=11(xiyi)iω(y+r(xy))rddrdy+di=1f(x)G(xiyi)2|xy|2ω(y)dy=Gdf(y)1ω(y+r(xy))rd1drdy+Gf(y)1ddrω(y+r(xy))rddrdy+f(x)=ω(x)Gf(y)dy+f(x)

Image

using (2.1.74), (2.1.75) and Gωdy=1Image. As Gfdy=0Image we obtain divu=fImage.

Now we pass to general functions fLA(G)Image. Recall that, if fC0,(G)Image, then uC0(G)Image, and moreover the equations (2.1.76) and (2.1.23) hold for every xGImage. Due to (2.0.11), LA(G)LLogL(G)Image, since B(t)Image grows at least linearly near infinity, and hence A(t)Image dominates the function tlog(1+t)Image near infinity. Since the space C0,(G)Image is dense in LlogL(G)Image, there exists a sequence of functions {fk}C0,(G)Image such that fkfImage in LlogL(G)Image. Hence

BogG:LlogL(G)L1(G)

Image

(in fact, BogGImage is also bounded into LlogL(G)Image). Furthermore,

Hij:LlogL(G)L1(G),

Image

as a consequence of (2.1.78) and of a special case of Theorem 2.1.9, with LA(G)=LlogL(G)Image and LB(G)=L1(G)Image. Thus, BogfkBogGfImage in L1(G)Image and HijfkHijfImage in L1(G)Image. This implies that uW1,10(G)Image, and (2.1.76) and (2.1.23) hold.

By Theorem 2.1.9, the singular integral operator defined by the first term on the right-hand side of (2.1.77) is bounded from LA(G)Image into LB(G)Image. By inequality (2.1.78), the operator defined by the second term on the right-hand-side of (2.1.77) has (at least) the same boundedness properties as a Riesz potential operator with kernel 1|xy|d1Image. Such an operator is bounded in L1(G)Image and in L(G)Image, with norms depending only on |G|Image and on d. An interpolation theorem by Calderon [19, Theorem 2.12, Chap. 3] then ensures that it is also bounded from LA(G)Image into LA(G)Image, and hence from LA(G)Image into LB(G)Image, with norm depending on d and |G|Image. Finally, the operator given by the last term on the right-hand-side of (2.1.77) is pointwise bounded (in absolute value) by |f(x)|Image. Thus, it is bounded from LA(G)Image into LA(G)Image, and hence from LA(G)Image into LB(G)Image. Equations (2.1.21) and (2.1.24) are thus established.

Inequality (2.1.25) follows from (2.1.24) via a scaling argument analogous to that which leads to (2.1.42) from (2.1.41) – see the Proof of Theorem 2.1.9.  □

Proof of Theorem 2.1.7

By an argument as in the proofs of [85, Lemmas 3.4 and 3.5, and Theorem 3.3], it suffices to show that, if G and G are bounded Lipschitz domains, for which the domain G0=GDImage is star-shaped with respect to a ball BG0Image, and f has the form

f=ζdivg+θGφdivgdy,

Image

for some functions ζC0(G)Image, θC0(G0)Image and φC(G)Image, and fulfills

G0f(x)dx=0,

Image

then there exists a function wW1,B0(G)Image such that

divw=finG0,

Image (2.1.79)

wLB(G0)CdivgLA(G),

Image (2.1.80)

and

wLB(G0)CgLA(G),

Image (2.1.81)

for some constant C=C(φ,θ,ζ,c,B,G,G)Image, where c is the constant appearing in (2.0.11) and (2.0.12).

Since fLA(G0)Image, an inspection of the proof of Theorem 2.1.6 then reveals that the function w, given by

w(x)=G0f(y)N(x,y)dyforxG0,

Image (2.1.82)

where

N(x,y)=xy|xy|d|xy|ω(y+ζxy|xy|)ζd1dζforx,yG,

Image (2.1.83)

and ω is any (nonnegative) function in C0(B)Image with Bωdx=1Image, satisfies (2.1.79), and

wLB(G0)CfLA(G0)

Image (2.1.84)

for some constant C. Since

fLA(G0)CdivgLA(G)

Image (2.1.85)

for some constant C, inequality (2.1.80) follows. It remains to prove (2.1.81). To this purpose, assume, for the time being, that divgC0(G)Image. Then, by [85, Equation 3.35],

wi(x)=G0Ni(x,y)g(y)ζ(y)dyG0Kij(x,xy)ζ(y)gj(y)dyG0nj=1Gij(x,y)ζ(y)gj(y)dyζ(x)dj=1gj(x)G0(xiyi)(xjyj)|xy|2ω(y)dy(Ggφdy)G0Ni(x,y)θ(y)dyfor a.e.xG0,

Image (2.1.86)

where the kernels KijImage and GijImage satisfy the same assumptions as the kernels in (2.1.77), and |N(x,y)|C|xy|1dImage for some constant C. Note that condition (2.1.28) has been used in writing the last term on the right-hand side of equation (2.1.86).

We now drop the assumption that divgC0(G)Image. Condition (2.0.11) entails that LA(G)LlogL(G)Image, and hence HA0(G)HLlogL0(G)Image, where the latter space denotes HA0(G)Image with A(t)Image equivalent to tlog(1+t)Image near infinity. Thus, gHLlogL0(G)Image, and hence it can be approximated by a sequence of functions {gk}C0(G)Image in such a way that gkgImage in HLlogL(G)Image, cf. Theorem A.1.46. The first and third term on the right-hand side of (2.1.86) are integral operators applied to g whose kernel is bounded by a multiple of |xy|1dImage. The fourth term is just bounded by a constant multiple of |g|Image. The last term is a constant multiple of an integral of g against a bounded vector-valued function ∇φ. Hence, all these operators are bounded from LA(G)Image into LA(G0)Image. The second term is a singular integral operator enjoying the same properties as the singular integral operator K in Theorems 2.1.8 and 2.1.9. Thus, since all operators appearing on the right-hand side of (2.1.86) are bounded from LlogL(G)Image into L1(G0)Image, the right-hand side of (2.1.86), evaluated with g replaced by gkImage, converges in L1(G0)Image to the right-hand-side of (2.1.86). On the other hand, equation (2.1.82) and the properties of the Bogovskiĭ  operator tell us that the left-hand-side of (2.1.86), with wiImage corresponding to gkImage, converges in L1(G0)Image to the left-hand-side of (2.1.86). Altogether, we conclude that (2.1.86) actually holds even if g is just in HA0(G)Image.

The properties of the operators on the right-hand-side of (2.1.86) mentioned above ensure that they are bounded from LA(G)Image into LB(G0)Image. Inequality (2.1.81) thus follows from (2.1.86).  □

2.2 Negative norms & the pressure

We need a last preliminary result in preparation of the proof of Theorem 2.0.5.

Proposition 2.2.1

Let G be an open subset in RdImage such that |G|<Image, and let A be a Young function. Assume that uLA(G)Image. Then we have

supvL˜A(G)GuvdxvL˜A(G)=supφC0(G)GuφdxφL˜A(G),

Image (2.2.87)

and

supvL˜A(G)GuvdxvL˜A(G)=supφC0,(G)GuφdxφL˜A(G).

Image (2.2.88)

Note that equation (2.2.87) is well known under the assumption that A2Image near infinity, namely ˜AΔ2Image near infinity. This is because C0(G)Image is dense in L˜A(G)Image in this case. Equation (2.2.88) also easily follows from this property when A2Image near infinity. The novelty of Proposition 2.2.1 is in the arbitrariness of A.

Proof of Proposition 2.2.1

Consider first (2.2.87). It clearly suffices to show that

supvL˜A(G)GuvdxvL˜A(G)=supvL(G)GuvdxvL˜A(G),

Image (2.2.89)

and

supvL(G)GuvdxvL˜A(G)=supφC0(G)GuφdxφL˜A(G).

Image (2.2.90)

Given any vL˜A(G)Image, define, for kNImage, the function vk:GRImage as

vk=sign(v)min{|v|,k}.

Image (2.2.91)

Clearly, vkL(G)Image, and 0|vk||v|Image a.e. in G as kImage. Hence,

G|uvk|dxG|uv|dxask,

Image

by the monotone convergence theorem for integrals, and, by the Fatou property of the Luxemburg norm,

vkL˜A(G)vL˜A(G)ask.

Image

Thus, since

supvL˜A(G)GuvdxvL˜A(G)=supvL˜A(G)G|uv|dxvL˜A(G),

Image

equation (2.2.89) follows.

As far as (2.2.90) is concerned, consider an increasing sequence of compact sets EkImage such that dist(Ek,RdG)2kImage, EkEk+1GImage for kNImage, and kEk=GImage. Moreover, let {ϱk}Image be a family of (nonnegative) smooth mollifiers in RdImage, such that suppϱkB1k(0)Image and Rdϱkdx=1Image for kNImage. Given vL(G)Image, define wk:RdRImage as

wk={vin Ek,0elsewhere,

Image

and φk:RdRImage as

φk(x)=Rdwk(y)ϱk(xy)dyforxRd.

Image (2.2.92)

Classical properties of mollifiers ensure that

φkC0(G),φkva.e. in G ask,φkL(G)vL(G)forkN.

Image

Thus, if uLA(G)Image, then

GuφkdxGuvdxask,

Image (2.2.93)

by the dominated convergence theorem for integrals. Moreover,

φkL˜A(G)vL˜A(G)ask.

Image (2.2.94)

Indeed, by dominated convergence and the definition of Luxemburg norm,

G˜A(|φk|vL˜A(G))dxG˜A(|v|vL˜A(G))dx1ask.

Image

In particular, for every ε>0Image, there exists kεImage such that

G˜A(|φk|vL˜A(G))dx<1+εifk>kε.

Image

Hence, by the arbitrariness of ε and the definition of Luxemburg norm,

liminfkφkL˜A(G)vL˜A(G).

Image (2.2.95)

We also have that

limsupkφkL˜A(G)vL˜A(G).

Image (2.2.96)

Indeed, assume that (2.2.96) fails. Then, there exists σ>0Image and a subsequence of {φk}Image, still denoted by {φk}Image, such that

1<G˜A(|φk|vL˜A(G)+σ)dxG˜A(|v|vL˜A(G)+σ)dx1,

Image

which is a contradiction. Equation (2.2.94) follows from (2.2.95) and (2.2.96). Coupling (2.2.93) with (2.2.94) implies (2.2.90). The proof of (2.2.87) is complete.

The proof of (2.2.88) follows along the same lines, and, in particular, via the equations

supvL˜A(G)GuvdxvL˜A(G)=supvL(G)GuvdxvL˜A(G),

Image (2.2.97)

and

supvL(G)GuvdxvL˜A(G)=supφC0,(G)GuφdxφL˜A(G).

Image (2.2.98)

For any vL˜A(G)Image, we define the sequence of functions {vk}L(G)Image by

vk=vk(vk)G.

Image

Here we have kNImage and vkImage is given by (2.2.91). We can prove equation (2.2.97) via a slight variant of the argument used for (2.2.89). Here, one has to use the fact that (vk)G0Image as kImage.

Equation (2.2.98) can be established similarly to (2.2.90). Let vL(G)Image be given. We have to replace the sequence {φk}Image defined by (2.2.92) with the sequence {φk}C0,(G)Image defined by

φk=φk(φk)GψforkN.

Image

Here ψ is any function in C0(G)Image such that Gψdx=1Image. For every ε>0Image, there exists kεNImage such that φkL(G)vL(G)+εImage, provided that k>kεImage.  □

Proof of Theorem 2.0.5

Let uL1(G)Image. Then

uuGLB(G)=supvL˜B(G)G(uuG)vdxvL˜B(G)=supvL˜B(G)G(uuG)(vvG)dxvL˜B(G)=supvL˜B(G)Gu(vvG)dxvL˜B(G)3supvL˜B(G)Gu(vvG)dxvvGL˜B(G)=3supvL˜B(G)GuvdxvL˜B(G)=3supφC0,(G)GuφdxφL˜B(G).

Image (2.2.99)

Note that the inequality in (2.2.99) holds since, by the first inequality in (1.2.3),

vvGL˜B(G)vL˜B(G)+vGL˜B(G)vL˜B(G)+|vG|1L˜B(G)vL˜B(G)+2|G|vL˜B(G)1LB(G)1L˜B(G)=vL˜B(G)+2|G|vL˜B(G)1B1(|G|)1˜B1(|G|)3vL˜B(G).

Image

The last equality in (2.2.99) follows from (2.2.88). By Theorem 2.1.6, applied with A and B replaced with ˜BImage and ˜AImage, respectively, there exists a constant C=C(G,c)Image such that

supφC0,(G)GuφdxφL˜B(G)=supφC0,(G)Gudiv(BogGφ)dxφL˜B(G)CsupφC0,(G)Gudiv(BogGφ)dxBogGφL˜A(G)CsupφC0(G)GudivφdxφL˜A(G)dx=CuW1,A(G).

Image (2.2.100)

The first inequality in (2.0.13) follows from (2.2.99) and (2.2.100). The second inequality is trivial, since

uW1,A(G)=supφC0(G)GudivφφL˜A(G)dx=supφC0(G)G(uuG)divφdxφL˜A(G)CsupφC0(G)G(uuG)divφdxdivφL˜A(G)dxCsupφC0,(G)G(uuG)φdxφL˜A(G)dx2CuuGLA(G),

Image

for some constant C=C(d)Image.  □

In the remaining part of this section, we focus on the second question raised in the introduction of this chapter, namely the reconstruction of the pressure π in a correct Orlicz space. In case of fluids governed by a general constitutive law, the function H belongs to some Orlicz space LA(G)Image. If AΔ22Image, then πLA(G)Image as well. However, in general, one can only expect that π belongs to some larger Orlicz space LB(G)Image. The balance between the Young functions A and B is determined by conditions (2.0.11) and (2.0.12), as stated in the following result.

Theorem 2.2.10

Let A and B be Young functions fulfilling (2.0.11) and (2.0.12). Let G be a bounded domain with the cone property in RdImage, d2Image. Assume that HLA(G)Image satisfies

GH:φdx=0

Image

for every φC0,div(G)Image. Then there exists a unique function πLB(G)Image such that

GH:φdx=Gπdivφdx

Image (2.2.101)

for every φC0(G)Image. Moreover, there exists a constant C=C(G,c)Image such that

πLB(G)CHHGLA(G),

Image (2.2.102)

and

GB(|π|)dxGA(C|HHG|)dx.

Image (2.2.103)

Here, c denotes the constant appearing in (2.0.11) and (2.0.12).

In particular, Theorem 2.2.10 reproduces, within a unified framework, various results appearing in the literature. For instance, when the power law model is in force, the function A(t)Image is just a power tqImage for some q>1Image. So LA(G)Image agrees with the Lebesgue space Lq(G)Image, and Theorem 2.2.10 recovers the fact that π belongs to the same Lebesgue space Lq(G)Image.

As far as the simplified system (without convective term) for the Eyring–Prandtl model (see Chapter 4) is concerned, under appropriate assumptions on the function f one has that H=S(ε(v))+Δ1fexpL(G)Image. Hence, via Theorem 2.2.10, we obtain the existence of a pressure πexpL12(G)Image. More generally, if HexpLβ(G)Image for some β>0Image, one has that πexpLβ/(β+1)(G)Image. The complete system for the Eyring–Prandtl model, in the 2-dimensional case, admits a weak solution v such that vvLlogL2(G)Image and hence H=S(ε(v))+Δ1fvvLlogL2(G)Image, see Chapter 4. Again, one cannot expect that the pressure π belongs to the same space. In fact, Theorem 2.2.10 implies the existence of a pressure πLlogL(G)Image. This reproduces a result from [33]. In general, if HLlogLα(G)Image for some α1Image, then we obtain that πLlogLα1(G)Image.

Proof of Theorem 2.2.10

By De Rahms Theorem, in the version of [131], there exists a distribution Ξ such that

GH:φdx=Ξ(divφ)

Image (2.2.104)

for every φC0(G)Image. Replacing φ with BogG(φφG)Image in (2.2.104), where φC0(G)Image, implies

GH:BogG(φφG)dx=Ξ(φφG)

Image

for every φC0(G)Image. We claim that the linear functional C0(G)φΞ(φφG)Image is bounded on C0(G)Image equipped with the L(G)Image norm. Indeed, by (2.0.11), one has that LA(G)LlogL(G)Image. Moreover, by a special case of Theorem 2.1.6, BogG:L(G)expL(G)Image. Thus, since LlogL(G)Image and expL(G)Image are Orlicz spaces generated by Young functions which are conjugate of each other,

|GH:BogG(φφG)dx|CHLlogL(G)BogG(φφ)G)expL(G)CHLA(G)φφGL(G)CHLA(G)φL(G),

Image (2.2.105)

for every φC0(G)Image, where C=C(|G|,d)Image and C=C(G,c)Image. Hence, the relevant functional can be continued to a bounded linear functional on φC0(G)Image, with the same norm.

Now, as a consequence of Riesz's representation Theorem, there exists a Radon measure μ such that

Ξ(φφG)=Gφdμ

Image

for every φC0(G)Image. Fix any open set EGImage. By Theorem 2.1.6 again, there exists a constant C such that

μ(E)=supφC00(E),φ=1Ξ(φφG)=supφC00(E),φ=1GH:BogG(φφG)dxsupφC00(E),φ=1HLlogL(E)BogG(φφG)expL(G)CsupφC00(E),φ=1HLlogL(E)φ(φ)GL(G)CHLlogL(E).

Image (2.2.106)

One can verify that the norm LlogL(E)Image is absolutely continuous in the following sense. For every ε>0Image there exists δ>0Image such that HLlogL(E)<εImage if |E|<δImage. Since any Lebesgue measurable set can be approximated from outside by open sets, inequality (2.2.106) implies that the measure μ is absolutely continuous with respect to the Lebesgue measure. Hence, μ has a density with respect to the Lebesgue measure. So Ξ can be represented by a function πL1(G)Image fulfilling (2.2.101) holds. The function π is uniquely determined if we assume that πG=0Image. By this assumption, Theorem 2.0.5, and equation (2.2.101) we have that

πLB(G)CπW1,A(G)=CsupφC0(G)GπdivφdxφL˜A(G)=CsupφC0(G)GH:φdxφL˜A(G)=CsupφC0(G)G(HHG):φdxφL˜A(G)2CHHGLA(G),

Image

where C=C(G,c)Image. This proves inequality (2.2.102). Inequality (2.2.103) follows from (2.2.102), by replacing A and B with kA and kB, respectively, with k=1GA(|HHG|)dxImage, via an argument analogous to that of the proof of (2.1.42).  □

2.3 Sharp conditions for Korn-type inequalities

In order to formulate the main result of this section we need to introduce the Banach function space

EA(G):={uLA(G):ε(u)LA(G)},uEA(G):=uLA(G)+ε(u)LA(G),

Image

and its subspace

EA0(G):={uEA(G): the continuation of u by zero belongs to EA(Rd)}.

Image

If A is of power growth (belongs to Δ22Image) then there is no need for this definition. In this case the space EA(G)Image coincides with the standard Lebesgue spaces (Orlicz spaces). However, this is not true in our general context.

Theorem 2.3.11

Let G be any open bounded set in RdImage. Let A and B be Young functions such that

ttt0B(s)s2dsA(ct)fortt0,

Image (2.3.107)

and

ttt0˜A(s)s2ds˜B(ct)fortt0,

Image (2.3.108)

for some constants c>0Image and t00Image. Then EA0(G)W1,B0(G)Image, and

uLB(G)Cε(u)LA(G)

Image (2.3.109)

for some constant C=C(t0,G,c,A,B)Image and for every uEA0(G)Image. Moreover,

GB(C|u|)dxC1+GA(C|ε(u)|)dx

Image

for every uEA0(G)Image with C1=C1(t0,G,c,A,B)Image. If t0=0Image then C1=0Image and C=C(G,c)Image.

Theorem 2.3.12

Let G be an open bounded Lipschitz domain in RdImage. Assume that A and B are Young functions fulfilling conditions (2.3.107) and (2.3.108). Then EA(G)W1,B(G)Image, and

u(u)GLB(G)Cε(u)(ε(u))GLA(G)

Image (2.3.110)

for some constant C=C(t0,G,c;A,B)Image and for every uEA(G)Image. Moreover,

GB(C|u(u)G|)dxC1+GA(C|ε(u)(ε(u))G|)dx

Image (2.3.111)

for every uEA(G)Image with C1=C1(t0,G,c,A,B)Image. If t0=0Image then C1=0Image and C=C(G,c)Image.

Instead of subtracting the mean value it is also useful to subtract an element from the kernel of the differential operator ε given by

R={w:RdRd:w(x)=b+Qx:bRd,QRd×d,Q=QT}.

Image

Corollary 2.3.1

Let G be an open bounded Lipschitz domain in RdImage. Assume that A and B are Young functions fulfilling conditions (2.3.107) and (2.3.108). Then EA(G)W1,B(G)Image, and there is wRImage such that

uwLB(G)Cε(u)LA(G)

Image (2.3.112)

for some constant C=C(t0,G,c;A,B)Image and for every uEA(G)Image. Moreover,

GB(C|uw|)dxC1+GA(C|ε(u)|)dx

Image (2.3.113)

for every uEA(G)Image with C1=C1(t0,G,c,A,B)Image. If t0=0Image then C1=0Image and C=C(G,c)Image.

Theorem 2.3.13

Let G be an open bounded Lipschitz domain in RdImage and let A be any Young function. Then there is wRImage such that

uwLA(G)Cε(u)LA(G)

Image (2.3.114)

for some constant C=C(G,A)Image and for every uEA(G)Image. Moreover,

GA(C|uw|)dxGA(C|ε(u)|)dx

Image (2.3.115)

for every uEA(G)Image.

Proof of Theorem 2.3.12

Let us introduce negative norms for single partial derivatives as follows. Given uL1(G)Image, we set

uxkW1,A(G)=supφC0(G)GuφxkdxφL˜A(G)fork=1,,d.

Image

Obviously, the following holds

uxkW1,A(G)uW1,A(G)fork=1,d.

Image (2.3.116)

On the other hand,

uW1,A(G)=supφC0(G)GudivφdxφL˜A(G)=supφC0(G)dk=1GuφkxkdxφL˜A(G)supφC0(G)dk=1GuφkxkdxφkL˜A(G)dk=1supφC0(G)GuφxkdxφL˜A(G)=dk=1uxkW1,A(G),

Image (2.3.117)

where φkImage denotes the k-th component of φ. Next, notice the identity

2vixkxj=(ε(v))ijxk+(ε(v))ikxj(ε(v))jkxi

Image (2.3.118)

for every weakly differentiable function v:GRdImage. Thus, the following chain holds for every uW1,1(G)EA(G)Image

u(u)GLB(G)Cdi,j=1uixj(uixj)GLB(G)Cni,j=1uixjW1,A(G)Cdi,j,k=12uixkxjW1,A(G)Cdi,j,k=1((ε(u))ijxkW1,A(G)+(ε(u))ikxjW1,A(G)+(ε(u))jkxiW1,A(G))Cdi,j=1(ε(u))ijW1,A(G)Cni,j=1(ε(u))ij((ε(u))ij)GLA(G)Cε(u)(ε(u))GLA(G).

Image (2.3.119)

Note that the second inequality holds by Theorem 2.0.5 (see Remark 2.0.6 for the case t0>0Image), the third by (2.3.117), the fourth by (2.3.118), the fifth by (2.3.116), and the sixth by Theorem 2.0.5 again.

Let us turn to the modular version. Suppose first that t0=0Image in (2.3.107) and (2.3.108). An inspection of the proof of Theorem (2.3.119) and of the statement of Theorem 2.0.5 tells us that the constant C in (2.3.119) depends only on G and on the constant c appearing in conditions (2.3.107) and (2.3.108). These conditions continue to hold if the functions A and B are replaced by the functions AMImage and BMImage given by AM(t)=A(t)/MImage and BM(t)=B(t)/MImage for some positive constant M. Given a function uEA0(G)Image, set

M=ΩA(C|ε(u)(ε(u))G|)dx.

Image

If M=Image, then inequality (2.3.111) holds trivially. We may thus assume that

u(u)GLAM(G)1.

Image

Hence, by inequality (2.3.110) applied with A and B replaced by AMImage and BMImage, we obtain

ΩB(|u(u)G|)dxΩA(C|ε(u)(ε(u))G|)dx.

Image (2.3.120)

This is (2.3.111) with C1=0Image.

Assume next that (2.3.107) and (2.3.108) just hold for some t0>0Image. The functions A and B can be replaced by new Young functions AImage and BImage, equivalent to A and B near infinity, and such that (2.3.107) and (2.3.108) hold for the new functions with t0=0Image. The same argument as above implies (2.3.120) with A and B replaced by AImage and BImage., i.e., we have

ΩB(|u(u)G|)dxΩA(C|ε(u)(ε(u))G|)dx

Image (2.3.121)

for some constant C. Since AImage and BImage are equivalent to A and B near infinity, there exist constants t0>0Image and c>0Image such that

A(t)A(ct)iftt0,B(t)B(ct)iftt0.

Image (2.3.122)

From (2.3.121) and (2.3.122) we obtain

GB(|u(u)G|)dx={|u(u)G|<t0}B(|u(u)G|)dx+{|u(u)G|t0}B(|u(u)G|)dxB(t0)|G|+GB(c|u(u)G|)dxB(t0)|G|+GA(Cc|ε(u)(ε(u))G|)dxB(t0)|G|+{Cc|ε(u)(ε(u))G|<t0}A(Cc|ε(u)(ε(u))G|)dx+{Cc|ε(u)(ε(u))G|t0}A(Cc|ε(u)(ε(u))G|)dx(B(t0)+A(ct0))|G|+GA(Cc2|ε(u)(ε(u))G|)dx,

Image

namely (2.3.111) with C1=(B(t0)+A(ct0))|G|Image.  □

Proof of Theorem 2.3.11

If uW1,10(G)Image, then (u)G=(ε(u))G=0Image, and Theorem 2.3.11 implies the claim.  □

Proof of Corollary 2.3.1

Let us choose wRImage through the condition w=(u)G(ε(u))GImage. Then the following holds by (2.3.110)

uwLB(G)u(u)GLB(G)+(ε(u))GLB(G)cε(u)LA(G).

Image

The integral version follows form the norm version as in the proof of (2.3.111).  □

Proof of Theorem 2.3.13

Step 1: Star-shaped domains.

First, we assume that G is star-shaped with respect to some ball BGImage. Then, according to formula (2.39) in [126] each vC(G)Image can be represented as

u(x)=PGu(x)+Lε(ε(u))(x),

Image (2.3.123)

where PG:L1(G)RImage is a suitable linear projection operator into the space of rigid motions (compare (2.33)–(2.39) in [126]) defined even for L1(G)Image functions. Furthermore, the operator LεImage is a weakly singular integral operator (compare (2.37) in [126]) given by

Lε(φ):=Sε(ψ)+Tε(ψ),ψC(G),Siε(φ)(x):=GGi(x,e)|xz|d1:ψ(z)dz,i=1,...,d,Tiε(φ)(x):=Gθi(x,z):ψ(z)dz,i=1,...,d,

Image (2.3.124)

for xGImage. Here, Gi(x,e)Image are smooth functions (e:=(xz)/|xz|Image) and θi(x,z)Image are bounded continuous functions (both with values in Rd×dImage); see [126] after (2.38).

Since the kernel of SεImage is essentially homogeneous of degree 1dImage the theory of Riesz potentials applies (see, e.g. [134] or [90]). Hence we have Sε:L1(G)L1(G)Image. Of course, the same is true for TεImage because the θi(x,z)Image are bounded. So, we obtain

Lε:L1(G)L1(G)

Image (2.3.125)

continuously. We claim that equation (2.3.123) continues to hold even if uEA(G)Image. From [138, Proposition 1.3, Chapter 1], we deduce that C(G)Image is dense in E1(G)Image. Thus, there exists a sequence {um}C(G)Image such that

umuinE1(G).

Image

We already know that formula (2.3.123) holds with u replaced by umImage. Thus we can pass to the limit (passing to a subsequence if necessary) in the representation formula (2.3.123) applied to umImage. This implies that it continues to hold also for u. Note that this is a consequence of (2.3.125).

Now, we turn to the case of a Lipschitz domain. In order to do so, we consider a general linear projection operator Π:L1(G)ΣImage such that

ΠuL1(G)CuL1(G),

Image (2.3.126)

for some constant C, and every uL1(G)Image. We claim that there exists a constant CImage such that

infwRuwLA(G)uΠuLA(G)C(A)infwRuwLA(G),

Image (2.3.127)

for every uLA(G)Image and for every Young function A.

The left-wing inequality in (2.3.127) is trivial. As far as the right-wing inequality is concerned, given any wRImage, and any u in LA(G)Image we set

v=w+(uw)G.

Image

Since Π, restricted to RImage, agrees with the identity map, we have that Πv=vImage. As a consequence,

uΠu=(uv)Π(uv).

Image

Thus,

uΠuLA(GuvLA(G)+Π(uv)LA(G).

Image (2.3.128)

By the triangle inequality,

uvLA(G)=uw(uw)GLA(G)2uwLA(G).

Image (2.3.129)

Since the range of Π is a finite dimensional space, where all norms are equivalent, there exists a constant CImage such that

Π(uv)LA(G)+Π(uv)LA(G)CΠ(uv)L1(G).

Image (2.3.130)

Inequality (2.3.126) ensures that

Π(uv)L1(G)CuvL1(G)=CuwuwGL1(G).

Image (2.3.131)

Now, by the triangle inequality,

uwuwGL1(G)2uwL1(G).

Image (2.3.132)

On the other hand, our assumptions on G ensure that a Poincaré-type inequality holds in W1,1(G)Image. Hence there exists a constant C such that

uwuwGL1(G)C(uw)L1(G).

Image (2.3.133)

Altogether, inequality (2.3.127) follows.

Step 2: the union of two sets.

Let A and B be Young functions. An open set G in RdImage, d2Image, will be called admissible with respect to the couple (A,B)Image if there exists a constant C such that

infwRuwLA(G)Cε(u)LA(G)

Image (2.3.134)

for every uEA(G)Image. We will show that, under certain assumptions, the union of two admissible sets will be admissible. So, assume that G1Image and G2Image are bounded connected open sets in RdImage with Lipschitz boundary. Assume that each of them is admissible with respect to (A,B)Image, and G1G2Image. Then we have

the set G1G2 is admissible with respect to (A,B).

Image (2.3.135)

In order to prove (2.3.135) we consider a ball BG1G2Image. Fix ωC0(B)Image. Denote by P1Image the space of polynomials of degree not exceeding 1, and by Π2uP1Image the averaged Taylor polynomial of second-order with respect to ω of a function uL1(G1G2)Image – see [37]. The operator Π2:L1(G1G2)P1Image is linear, and, by [37, Corollary 4.1.5], there exists a constant C such that

Π2uL1(G1G2)CuL1(B)

Image

for every uL1(G1G2)Image. Let us denote by ΠRImage the L2Image-orthogonal projection from P1Image into RImage. We have that

ΠRpL1(G1G2)cpL1(G1G2)

Image

for every pP1Image. Thus, the linear operator Π=ΠRΠ2Image maps L1(G1G2)Image into RImage, and there exists a constant C such that

ΠuL1(Gj)ΠuL1(G1G2)CuL1(B)CuL1(Gj)j=1,2

Image (2.3.136)

for every uL1(G1G2)Image. Due to inequality (2.3.136), inequality (2.3.127) ensures that there exists a constant C such that

infwRuwLA(G1G2)uΠuLA(G1G2)j=1,2uΠuLA(Gj)Cj=1,2infwRuwLA(Gj)

Image (2.3.137)

for every uLA(G)Image. The conclusion (2.3.135) follows from (2.3.137) and (2.3.134) applied with G=GjImage, for j=1,2Image.

Step 3: The general cases.

Any open Lipschitz domain G is the finite union of open sets GiImage, i=1,,kImage, starshaped with respect to a ball. Since G is connected, after, possibly, relabelling, we may assume that the sets j1i=1GiImage and GjImage have a non-empty intersection. The conclusion then follows from repeated use of (2.3.135).  □

Remark 2.3.7

A proof of Corollary 2.3.1 can also be given based on the representation formula (2.3.123). After differentiating (2.3.123) the main part is a singular operator. It can be shown that it enjoys the properties (2.1.34)(2.1.39). Hence it is continuous from LA(G)Image to LB(G)Image thanks to Theorem 2.1.9. We refer to [31] and [46] for details.

In the following we are concerned with the necessity of the conditions (2.3.107) and (2.3.108) for a Korn-type inequality.

Theorem 2.3.14

Let G be an open bounded set in RdImage. Let A and B be Young functions such that

uLB(G)Cε(u)LA(G)

Image (2.3.138)

for some constant C and for every uW1,10(G)EA0(G)Image. Then conditions (2.3.107) and (2.3.108) hold.

Our proof of inequality (2.3.107) is based on the technique of laminates as in [52]. In general a first order laminate is a probability measure ν on Rd×dImage given by

ν=λδA+(1λ)δB

Image

where λ(0,1)Image and A,BRd×dImage with rank(AB)=1Image. Here δXImage denotes the Dirac measure concentrated on the matrix X. We say ν has average C if λA+(1λ)B=CImage. We obtain a second order laminate if we replace δAImage (resp. δBImage) by a first order laminate with average A (resp. B). Iteratively we can define laminates of arbitrary order with a given average. For a detailed discussion we refer to [102] and [118].

Lemma 2.3.1

[52], eq. (5)

Let ν be a laminate with average C, then there is a sequence of uniformly Lipschitz continuous functions ui:(0,r)dRdImage with boundary data Cx such that

(0,r)dΦ(|ui|)dxrdRd×dΦ(|X|)dν(X),

Image (2.3.139)

for every continuous function Φ.

Proof of Theorem 2.3.14

Part 1: Inequality (2.3.108) holds.

Assume, without loss of generality, that the unit ball B1Image, centered at 0, is contained in G, and denote by ωdImage its Lebesgue measure. Let us preliminarily observe that inequality (2.3.138) implies that

A dominates B near infinity.

Image (2.3.140)

Indeed, given any nonnegative function hLA(0,ωd)Image, consider the function v:B1RdImage given by

v(x)=(1|x|h(ωdrd)dr,0,,0)forxB1.

Image

Then vLA(B1)Image, and

|ε(v)(x)||v(x)|=h(ωd|x|d)forxB1.

Image

An application of (2.3.138), with u replaced by v, shows that

hLB(0,ωd)=vLB(Ω)Cε(v)LA(Ω)CvLA(Ω)=hLA(0,ωd).

Image

Thus LA(0,ωd)LB(0,ωd)Image, and (2.3.140) follows.

Now, given h as above, define the function ρ:[0,1][0,]Image by

ρ(r)=1rh(ωdtd)tdtforr[0,1],

Image

and the function v:B1RdImage by

u(x)=Qxρ(|x|)forxB1,

Image

where QRd×dImage is any skew-symmetric matrix such that |Q|=1Image. We have that u is a weakly differentiable function, and

ε(u)(x)=Qxx|x|2ρ(|x|)|x|,u(x)=Qρ(|x|)+Qxx|x|2ρ(|x|)|x|

Image

for a.e. xB1Image. Here, ⊙ denotes the symmetric part of the tensor product of two vectors in RdImage. Hence,

|ε(u)(x)||ρ(|x|)||x|=h(ωd|x|),ρ(|x|)|u(x)|+|ρ(|x|)||x|=|u(x)|+h(ωd|x|)

Image

for a.e. xB1Image. Thus, due to (2.3.138) and (2.3.140),

ωdsh(r)rdrLB(0,ωd)=1|x|h(ωdtd)tdtLB(B1)=ρ(|x|)LB(B1)uLB(B1)+h(ωd|x|d)LB(B1)Cε(u)LA(B1)+h(ωd|x|d)LA(B1)Ch(ωd|x|d)LA(B1)=Ch(s)LA(0,ωd)

Image (2.3.141)

for suitable constants C and CImage. Thanks to the arbitrariness of h, inequality (2.3.141) implies, via Lemma 1.2.1, that (5.3.12) holds for some c and t0Image.

Part 2: Inequality (2.3.107) holds.

Let us preliminarily note that, if A(t)=Image for large t, then (2.3.107) holds trivially. We may thus assume that A is finite-valued, and hence continuous. By (2.3.140), the function B is also finite-valued and continuous.

For ease of notations, we hereafter focus on case when d=2Image. An analogous argument carries over to any dimension along the lines of [52, Lemma 3]. Given a,bRImage, we define the matrix Ga,bImage as

Ga,b=(0ab0),

Image

and set δa,b=δGa,bImage. Next, we define the sequence {μ(m)}Image of laminates of order 2m iteratively by

{μ(0)=δt,t,μ(m)=13δ2mt,2mt+16δ21mt,21mt+12μ(m1)

Image (2.3.142)

for mNImage. We claim that μ(m)Image is a laminate with average G2mt,2mtImage for mNImage. Indeed, the following holds

μ(m)=14δ21mt,21mt+34μ(m1).

Image (2.3.143)

Since rank(Gt,tGt,t)=1Image, the right-hand side of (2.3.143) is a laminate with average G21t,tImage for m=1Image. Hence, μ(1)Image is a laminate with average G21t,21tImage. An induction argument then proves our claim. Now, note the representation formula

μ(m)=2mδt,t+mk=1(132kmδ2kt,2kt+162kmδ21kt,21kt)

Image (2.3.144)

for mNImage. We remark that δt,tImage is concentrated at a symmetric matrix, whereas the sum in (2.3.144) is concentrated at skew-symmetric matrices. We define the functions Φj:R2×2[0,)Image, for j=1,2Image, by

Φ1(X)=A(|XsymG2mt,2mt|),Φ2(X)=B(C1|XG2mt,2mt|),

Image

for XR2×2Image. Here, Xsym=12(X+XT)Image is the symmetric part of X, and C is the constant appearing in (2.3.138). Fix mNImage. Without loss of generality, we may assume that 0GImage. We choose r>0Image so small that (0,r)2GImage. Let mNImage be arbitrary. Due to Lemma 2.3.1 applied with ν=μ(m)Image, there exists a sequence {ui}Image of Lipschitz continuous functions ui:(0,r)2R2Image, such that ui(x)=G2mt,2mtxImage on (0,r)2Image, and

limi(0,r)2Φj(ui)dx=r2R2×2Φj(X)dμ(m)(X)forj=1,2.

Image (2.3.145)

We define the sequence {vi}Image of functions vi:GRImage by vi(x)=ui(x)G2mt,2mtxImage if x(0,r)2Image, and vi(x)=0Image if G(0,r)2Image. Then viW1,0(G)Image, and, by (2.3.145),

limiGA(|ε(vi)|)dx=limi(0,r)2A(|ε(v)i|)dx

Image (2.3.146)

=r2R2×2A(|(XsymG2mt,2mt)|)dμ(m)(X),limiΩB(C1|vi|)dx=limi(0,r)2B(C1|vi|)dx=r2R2×2B(C1|XG2mt,2mt|)dμ(m)(X).

Image (2.3.147)

We have the following chain

R2×2A(|XsymG2mt,2mt|)dμ(m)(X)12R2×2A(2|Xsym|)dμ(m)(X)+12R2×2A(2|G2mt,2mt|)dμ(m)(X)=122mA(2|Gt,t|)+12A(2|G2mt,2mt|)=122mA(2|Gt,t|)+12A(22m|Gt,t|)2mA(2|Gt,t|).

Image (2.3.148)

Here the first inequality holds since A is convex, the first equality holds due to (2.3.144) and to the fact that μ(m)Image is a probability measure, and the last inequality follows from (1.2.6). Combining (2.3.146) with (2.3.148) implies

limiGA(|ε(vi)|)dxr22mA(2|Gt,t|).

Image (2.3.149)

Since A is a continuous function, there exists tm(0,)Image such that

r22mA(2|Gtm,tm|)=12.

Image (2.3.150)

Thanks to (1.2.6), there exists t0>0Image, independent of m, such that

tmt02m.

Image (2.3.151)

Therefore, by neglecting, if necessary, a finite number of terms of the sequence {vi}Image, we can assume that

GA(|ε(vi)|)dx1

Image

for iNImage. Hence, ε(vi)A1Image for iNImage, and, by (2.3.138), viBCImage for iNImage. Thus,

GB(C1|vi|)dx1

Image

for iNImage. Combining the latter inequality with equation (2.3.147) implies us that

r2R2×2B(C1|XG2mtm,2mtm|)dμ(m)(X)1.

Image (2.3.152)

Next, one can make use of (2.3.144) and obtain the following chain

r2R2×2B(C1|XG2mtm,2mtm|)dμ(m)(X)2mB(C1(12m)|Gtm,tm|)+mk=1132kmB(C1(2k2m)|Gtm,tm|)+mk=1162kmB(C1(21k2m)|Gtm,tm|)m1k=1132kmB(C1(2k2m)|Gtm,tm|)m1k=1132kmB(C12k1|Gtm,tm|)m1k=11312C2mtmB(12C2ktm)12C2ktm.

Image (2.3.153)

It follows from (2.3.150), (2.3.152) and (2.3.153) that

22mA(2|Gtm,tm|)m1k=11312C2mtmB(12C2ktm)12C2ktm.

Image

Hence, by (2.3.151),

A(ctm)ctmm1k=1B(12C2ktm)12C2ktmctmtm4C2mtm4CB(s)s2dsctmtm4Ct02CB(s)s2ds

Image (2.3.154)

for suitable positive constants c, cImage cImage. Since limmtm=Image, one can find ˆtt02CImage such that, if t>ˆtImage, then there exists mNImage such that tmt<tm+1Image. Moreover, ˆtImage can be chosen so large that A is invertible on [ˆt,)Image and

tm=c1A1(c22m)

Image

for some positive constants c1,c2Image. By (1.2.7), the latter equation ensures that tm+12tmImage for mNImage. Thus, due to inequality (2.3.154),

A(2ct)A(2ctm)A(ctm+1)ctm+1tm+14Ct02CB(s)s2dsctt4Ct02CB(s)s2ds

Image

for tˆtImage. Hence, inequality (2.3.107) follows for suitable constants c and t0Image.  □

Corollary 2.3.2

Let G be a Lipschitz domain in RdImage, d2Image. Let A and B be Young functions. Assume that there exists a constant C such that

uuGLB(G)CuW1,A(G)

Image (2.3.155)

for every uL1(G)Image. Then conditions (2.3.107) and (2.3.108) hold.

Proof

Let uW1,A0(G)Image. As in the proof of Theorem (2.3.119) we can show the following chain

u(u)GLB(G)=u(u)GLB(G)Cdi,j=1uixj(uixj)GLB(G)Cni,j=1uixjW1,A(G)Cdi,j,k=12uixkxjW1,A(G)Cdi,j,k=1((ε(u))ijxkW1,A(G)+(ε(u))ikxjW1,A(G)+(ε(u))jkxiW1,A(G))Cdi,j=1(ε(u))ijW1,A(G)Cni,j=1(ε(u))ij((ε(u))ij)GLA(G)Cε(u)(ε(u))GLA(G)=Cε(u)LA(G).

Image

Here the second inequality in line two holds by (2.3.155). The first inequality in line four is true for every Young function by the very definition of the negative norm. The conclusion follows via Theorem 2.3.14, due to the arbitrariness of u.  □

Corollary 2.3.3

Let G be a bounded domain in RdImage, d2Image, which is starshaped with respect to a ball, and let BogGImage be the Bogovskiĭ  operator on G (see Section 2.1). Let A and B be Young functions such that

BogGfLB(G,Rd)CfLA(G)

Image (2.3.156)

for some constant C, and for every fC0,(G)Image. Then conditions (2.0.11) and (2.0.12) hold.

Proof

A close inspection of the proof of Theorem 2.0.5 reveals that inequality (2.3.156) implies inequality (2.3.155). The conclusion thus follows from Corollary 2.3.2.  □

References

[4] E. Acerbi, G. Mingione, Regularity results for stationary electro-rheological fluids, Arch. Ration. Mech. Anal. 2002;164:213–259.

[5] R.A. Adams, Sobolev Spaces. Pure Appl. Math.. New York: Academic Press, Inc.; 1975;vol. 65.

[15] R.J. Bagby, D.S. Kurtz, A rearranged good λ inequality, Trans. Am. Math. Soc. 1986;293:71–81.

[18] C. Bennett, K. Rudnick, On Lorentz–Zygmund spaces, Diss. Math. 1980;185.

[19] C. Bennett, R. Sharpley, Interpolation of Operators. Boston: Academic Press; 1988.

[22] M. Bildhauer, M. Fuchs, Variants of the Stokes problem: the case of anisotropic potentials, J. Math. Fluid Mech. 2003;5:364–402.

[24] M.E. Bogovskiĭ, Solutions of some problems of vector analysis, associated with the operators div and grad, In: Theory of Cubature Formulas and the Application of Functional Analysis to Problems of Mathematical Physics. Novosibirsk: Akad. Nauk SSSR Sibirsk. Otdel. Inst. Mat.; 1980:5–40 149 (in Russian).

[25] D. Breit, Analysis of generalized Navier–Stokes equations for stationary shear thickening flows, Nonlinear Anal. 2012;75:5549–5560.

[31] D. Breit, A. Cianchi, L. Diening, Trace-free Korn inequalities in Orlicz spaces, preprint at arXiv:1605.01006v1.

[32] D. Breit, L. Diening, Sharp conditions for Korn inequalities in Orlicz spaces, J. Math. Fluid Mech. 2012;14:565–573.

[33] D. Breit, L. Diening, M. Fuchs, Solenoidal Lipschitz truncation and applications in fluid mechanics, J. Differ. Equ. 2012;253:1910–1942.

[35] D. Breit, M. Fuchs, The nonlinear Stokes problem with general potentials having superquadratic growth, J. Math. Fluid Mech. 2011;13:371–385.

[37] S.C. Brenner, L.R. Scott, The Mathematical Theory of Finite Element Methods. Texts Appl. Math.. New York: Springer-Verlag; 1994;vol. 15.

[46] A. Cianchi, A sharp embedding theorem for Orlicz–Sobolev spaces, Indiana Univ. Math. J. 1996;45:39–65.

[47] A. Cianchi, Strong and weak type inequalities for some classical operators in Orlicz spaces, J. Lond. Math. Soc. 1999;60:187–202.

[48] A. Cianchi, Korn type inequalities in Orlicz spaces, J. Funct. Anal. 2014;267:2313–2352.

[51] R.R. Coifman, C. Fefferman, Weighted norm inequalities for maximal functions and singular integrals, Stud. Math. 1974;51:241–250.

[52] S. Conti, D. Faraco, F. Maggi, A new approach to counterexamples to L1 estimates: Korn's inequality, geometric rigidity, and regularity for gradients of separately convex functions, Arch. Ration. Mech. Anal. 2005;175:287–300.

[53] R. Courant, D. Hilbert, Methoden der mathematischen Physik, vol. 2. Berlin: Springer; 1937.

[59] L. Diening, P. Kaplicky, LqImage theory for a generalized Stokes System, Manuscr. Math. 2013;141:333–361.

[64] L. Diening, M. Růžička, K. Schumacher, A decomposition technique for John domains, Ann. Acad. Sci. Fenn., Math. 2010;35:87–114.

[66] G. Duvaut, J.L. Lions, Inequalities in Mechanics and Physics. Springer Grundlehren. Berlin: Springer; 1976;vol. 219.

[69] H.J. Eyring, Viscosity, plasticity, and diffusion as example of absolute reaction rates, J. Chem. Phys. 1936;4:283–291.

[70] D.M. Èidus, On a mixed problem of the theory of elasticity, Dokl. Akad. Nauk SSSR 1951;76:181–184.

[74] M. Fuchs, On stationary incompressible Norton fluids and some extensions of Korn's inequality, Z. Anal. Anwend. 1994;13(2):191–197.

[75] M. Fuchs, Korn inequalities in Orlicz spaces, Ir. Math. Soc. Bull. 2010;65:5–9.

[84] K. Friedrichs, On the boundary value problems of the theory of elasticity and Korn's inequality, Ann. Math. 1947;48(2):441–471.

[85] G.P. Galdi, An introduction to the mathematical theory of the Navier–Stokes equations, vol. I. Springer Tracts Nat. Philos.. Berlin–New York: Springer; 1994;vol. 38.

[90] D. Gilbarg, N.S. Trudinger, Elliptic Partial Differential Equations of Second Order. 2nd edition, revised 3rd printing Grundlehren Math. Wiss.. Berlin–Heidelberg–New York: Springer; 1998;vol. 224.

[91] J. Gobert, Une inéquation fondamentale de la théorie de l'élasticité, Bull. Soc. R. Sci. Liège 1962;3–4:182–191.

[92] J. Gobert, Sur une inégalité de coercivité, J. Math. Anal. Appl. 1971;36:518–528.

[102] B. Kirchheim, S. Müller, V. Svérak, Studying nonlinear pde by geometry in matrix space, In: S. Hildebrandt, H. Karcher, eds. Geometric Analysis and Nonlinear Partial Differential Equations. Springer; 2003:347–395.

[103] V. Kokilashvili, M. Krbec, Weighted Inequalities in Lorentz and Orlicz Spaces. Singapore: World Scientific Publishing; 1991.

[104] A. Korn, Über einige Ungleichungen, welche in der Theorie der elastischen und elektrischen Schwingungen eine Rolle spielen, Classe des Sciences Mathématiques et Naturels. 9, Novembre. Bull. Internat. Acad. Sci. Cracovie 1909:705–724.

[111] J. Málek, J. Necǎs, M. Rokyta, M. Růžička, Weak and Measure Valued Solutions to Evolutionary PDEs. London–Weinheim–New York: Chapman & Hall; 1996.

[114] R. von Mises, Mechanik der festen Körper in plastisch-deformablem Zustand, Götti. Nachr., Math.-Phys. Klasse 1913:582–592 Berlin.

[116] P.P. Mosolov, V.P. Mjasnikov, On the correctness of boundary value problems in the mechanics of continuous media, Math. USSR Sb. 1972;17(2):257–267.

[118] S. Müller, Variational models for microstructure and phase transitions, In: Calculus of Variations and Geometric Evolution Problems. F. Bethuel, ed. Springer Lect. Notes Math.. Berlin: Springer; 1999;vol. 1713:85–210.

[119] J. Nečas, Sur les normes équivalentes dans W(k)p(Ω)Image et sur la coecivité des formes formellement positives, In: Séminaire des mathématiques supérieures. Equ. Dériv. Partielles. Montréal: Les Presses de l'Université de Montréal; 1965;vol. 19:102–128.

[120] D. Ornstein, A non-inequality for differential operators in the L1Image norm, Arch. Ration. Mech. Anal. 1964;11:40–49.

[126] Y.G. Reshetnyak, Estimates for certain differential operators with finite-dimensional kernel, Sib. Math. J. 1970;11(2):315–326.

[131] J. Simon, Démonstration constructive d'un théorème de G. de Rahm, C. R. Acad. Sci. Paris 1993;316:1167–1172.

[134] E.M. Stein, Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, In: Monographs in Harmonic Analysis, III. Princeton, NJ: Princeton University Press; 1993 with the assistance of Timothy S. Murphy.

[138] R. Teman, Problèmes mathémathiques en plasticité. Paris: Gaulliers-Villars; 1985.

[143] E. Zeidler, Nonlinear Functional Analysis II/B–Nonlinear Monotone Operators, vol. 120. Berlin–Heidelberg–New York: Springer Verlag; 1990.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.188.57.172