• 沒有找到結果。

A Div-Curl Decomposition for the Local Hardy Space

N/A
N/A
Protected

Academic year: 2021

Share "A Div-Curl Decomposition for the Local Hardy Space"

Copied!
11
0
0

加載中.... (立即查看全文)

全文

(1)

A div-curl decomposition for the local

Hardy space

(to appear in the Proceedings of the American Mathematical Society)

Der-Chen Chang, Georgetown University, Department of Mathematics, Washington, DC 20057, USA

,

chang@georgetown.edu∗

Galia Dafni, Concordia University, Department of Mathematics & Statistics, Montreal, Quebec, H3G-1M8, CANADA

,

gdafni@mathstat.concordia.ca† Hong Yue, Trine University, Department of Mathematics & Informatics,

Angola, IN 46703, USA

,

yueh@trine.edu‡

June 11, 2009

Abstract

A decomposition theorem for the local Hardy space of Goldberg, in terms of nonhomogeneous div-curl quantities, is proved via a dual result for the space bmo.

1

Introduction

Given vector fields V = (v1, . . . , vn) in Lp(Rn, Rn), W = (w1, . . . , wn) in

Lp0

(Rn

, Rn) with 1 < p < ∞, 1 p +

1

p0 = 1, the scalar (dot) product V · W =

P viwiwill lie in L1(Rn). Coifman, Lions, Meyer, and Semmes [CLMS] showed

that if the following conditions hold in the sense of distributions:

div V :=X∂vi ∂xi = 0, curl W := ∂wj ∂xi −∂wi ∂xj  ij = 0,

Partially supported by a Hong Kong RGC competitive earmarked research grant #600607

and a competitive research grant at Georgetown University.

Partially supported by the Natural Sciences and Engineering Research Council, Canada.Partially supported by the Natural Sciences and Engineering Research Council,

(2)

then V · W belongs to the real Hardy space H1(Rn), a proper subspace of L1. Moreover,

kV · WkH1 ≤ CkVkLpkWkLp0. (1.1)

This “div-curl lemma” and other results of a similar nature illustrate the re-cent use of Hardy spaces for applications to nonlinear PDE and in the method of compensated compactness, originally going back to the work of Murat and Tartar.

Recall (see [FS]) that a function f belongs to H1

(Rn) if the maximal function

Mϕ(f ) belongs to L1(Rn), where, for a fixed Schwartz function ϕ, R ϕ = 1, we

define

Mϕ(f )(x) = sup t>0

|f ∗ ϕt(x)|, ϕt(·) = t−nϕ(t−1·).

Here the norm, defined by kf kH1 := kMϕkL1, depends on the choice of ϕ, but

the space does not since different choices of ϕ give equivalent norms. Functions in the Hardy space enjoy both improved integrability and cancellation conditions compared to L1functions. In particular, if f ∈ H1then its integral must vanish.

The local real Hardy space h1

(Rn), defined by Goldberg [Go], is larger than

H1 and allows for more flexibility, since global cancellation conditions are not

necessary. For example, the Schwartz space is contained in h1 but not in H1,

and multiplication by cut-off functions preserves h1but not H1, thus making it

more suitable for working in domains and on manifolds. For membership of a function f in h1(Rn), we use a “local” maximal function mϕ(f ), with

mϕ(f )(x) = sup 0<t<1

|f ∗ ϕt(x)|,

and require mϕ(f ) ∈ L1(Rn). As for H1, kf kh1 := kmϕkL1 defines a norm, and

different choices of ϕ give equivalent norms, so we can choose ϕ with compact support, making clear the local nature of the maximal function.

In [D], sufficient nonhomogeneous conditions on the divergence and curl were found in order for the dot product V · W, where V and W are vector fields as above, to be in h1

(Rn). In particular, as a corollary, the following special case

was proved:

Proposition 1.1 ([D]) Suppose V and W are vector fields on Rn satisfying

V ∈ Lp(Rn)n, W ∈ Lp0(Rn)n, 1 < p < ∞, 1 p+

1 p0 = 1.

If there exists a function f in Lp(Rn) and a matrix-valued function A with components in Lp0(Rn) such that, in the sense of distributions,

div V = f, curl W = A,

then V · W belongs to the local Hardy space h1

(Rn) with

kV · Wkh1≤ C (kVkLpkWkLp0 + kf kLpkWkLp0+ kVkLp

X

i,j

(3)

Note that the requirement on the divergence and the components of the curl to be functions in Lp and Lp0, respectively, was also used in the original div-curl lemma of Murat [Mu], and is a natural relaxation of the vanishing divergence and curl conditions. In fact, by the Hodge decomposition, this can be viewed as a combination of the following two theorems:

Theorem 1.2 ([D]) Suppose V and W are vector fields on Rn satisfying

V ∈ Lp(Rn)n, W ∈ Lp0(Rn)n, 1 < p < ∞, 1 p+ 1 p0 = 1 and div V = f ∈ Lp(Rn), curl W = 0

in the sense of distributions. Then V·W belongs to the local Hardy space h1

(Rn)

with

kV · Wkh1(Rn)≤ C (kVkLp(Rn)+ kf kLp(Rn)) kWkLp0(Rn); (1.2)

and

Theorem 1.3 ([D]) Suppose V and W are vector fields on Rn satisfying

V ∈ Lp(Rn)n, W ∈ Lp0(Rn)n, 1 < p < ∞, 1 p+

1 p0 = 1.

If A is a matrix with components in Lp0

(Rn) and

div V = 0, curl W = A

in the sense of distributions, then V · W belongs to the local Hardy space h1

(Rn) with kV · Wkh1(Rn)≤ C kVkLp(Rn) h kWkLp0(Rn)+ X i,j kAijkLp0(Rn) i . (1.3)

It may seem that the conditions on the divergence and the curl in the theo-rems above are too strong, so the question arises as to whether one can charac-terize functions in the Hardy space in terms of such div-curl quantities. Such a characterization, providing a kind of converse to the div-curl lemma, was shown in [CLMS] for H1(Rn):

Theorem 1.4 ([CLMS]) Every function f ∈ H1(Rn) can be written as

f =

X

k=1

λkVk· Wk,

with {λk} ∈ `1and Vk, Wk vector fields with norm bounded by 1 in L2(Rn, Rn),

(4)

This decomposition was proved in [CLMS], via functional analysis argu-ments, from the following dual result: for g ∈ BMO(Rn),

kgkBMO≈ sup

V,W

Z

Rn

gV · W, (1.4)

where the supremum is taken over all vector fields V, W in L2

(Rn

, Rn), kVk

L2, kWkL2 ≤

1, satisfying div V = 0, curl W = 0 in the sense of distributions on Rn.

The goal of this paper is to prove an analogue, Theorem 2.2, of (1.4) for functions in bmo(Rn), the dual of the local Hardy space h1(Rn), and conse-quently a decomposition theorem, Theorem 3.1, for h1in terms of the div-curl quantities used in Theorem 1.2 or in Theorem 1.3.

2

The div-curl lemma for local BMO

In [Go] it was shown that the dual of h1(Rn) can be identified with the space bmo(Rn), consisting of locally integrable functions f with

kf kbmo:= sup |I|≤1 1 |I| Z I |f − fI| + sup |I|>1 1 |I| Z I |f | < ∞.

Here the supremum can be taken over balls or cubes with sides parallel to the axes, |I| denotes Lebesgue measure (volume) and fI is the mean of f over I,

i.e. |I|1 R

If . The number 1 in the definition can be replaced by any other finite

nonzero constant. Note that unlike the case of BMO, we do not need to consider this norm modulo constants.

Before stating the main result, let us introduce the following definition in order to simplify notation.

Definition 2.1 Denote by DC1,0p the collection of all functions which can be written in the form V·W, where V, W are vectors fields, V ∈ Lp

(Rn

, Rn), W ∈

Lp0

(Rn

, Rn), kVk

Lp≤ 1, kWkLp0 ≤ 1, satisfying, in the sense of distributions

on Rn,

div V = f ∈ Lp(Rn), kf kLp ≤ 1, and curl W = 0. (2.1)

The collection DC0,1p can be defined analogously by requiring the vector fields to satisfy, instead of (2.1), the conditions

div V = 0, (curl W)ij = Aij∈ Lp

0

(Rn), kAijkLp0 ≤ 1, i, j ∈ {1, . . . , n}. (2.2)

Note that by Theorems 1.2 and 1.3, both DC1,0p and DC0,1p are subsets of h1(Rn) for every 1 < p < ∞.

(5)

Theorem 2.2 If g ∈ bmo(Rn), then for 1 < p < ∞, kgkbmo≈ sup f ∈DC1,0p Z Rn gf, (2.3) and kgkbmo≈ sup f ∈DC0,1p Z Rn gf, (2.4)

with constants that depend only on p and the dimension.

Proof: Let g ∈ bmo(Rn) (we are assuming that g is real-valued) and take

f ∈ DC1,0p ∪ DC0,1p . As stated above, by Theorems 1.2 and 1.3, f belongs to h1

(Rn

) with norm bounded by a constant. The duality of bmo(Rn) with h1

(Rn)

then gives

Z

Rn

gf ≤ Ckgkbmo.

It remains to prove the other direction, i.e.

kgkbmo ≤ C0

Z

Rn

gf

where we take the supremum over f ∈ DC1,0p or f ∈ DC0,1p , respectively. We will use the definition of bmo with the supremum taken over balls. Case I: If B is a small ball, say with radius bounded by 1 (although in fact the proof is independent of the radius), one can use the following estimate from the proof of Theorem III.2 in [CLMS]:

 1 |B| Z B g(x) − gB 2 dx 1/2 ≤ Cn sup Z g V · W, (2.5)

where the supremum is taken over all vector fields V, W ∈ C0∞( eB) (here and below eB denotes the ball with the same center and twice the radius of B) with kVkLp ≤ 1, kWkLp0 ≤ 1 and div V = 0, curl W = 0. Note that the result in

[CLMS] is stated for a cube Q (instead of a ball B) and for p = p0= 2, but can be modified to the case of a ball and for p 6= 2 as follows.

The key inequality used in the proof in [CLMS],

kg − gBkL2(B)≤ C n X i=1 ∂g ∂xi W−1,2(B) , (2.6)

is valid for any Lipschitz domain (see for example [GR], Section I.2.1, Corollary 2.1). This inequality holds in the homogeneous sense, modulo constants. There-fore, while we denote by W−1,2(B) the dual of the Sobolev space W01,2(B), which is the closure of C0∞(B) under the norm kϕkW1,2(B)= kϕkL2(B)+k∇ϕkL2(B), to

obtain the right-hand-side of (2.6) we test against test functions only bounded in the homogeneous Sobolev norm kukW˙1,2(B)= k∇ukL2(B). For a fixed ball the

(6)

homogeneous and nonhomogeneous norms on C0∞(B) functions are equivalent, but here we use only the homogeneous norm in order for the constant to be independent of the radius of the ball.

To estimate ∂g ∂xi

W−1,2(B), i = 1, . . . , n, we fix a j ∈ {1, . . . , n} \ {i} and

define the vector fields V and W, for the case p ≤ 2, as in [CLMS]: given u ∈ C0∞(B) with k∇ukL2≤ 1, let

V = |B|1/2−1/p ∂u ∂xi ej− ∂u ∂xj ei  , (2.7)

where ei denotes the ith coordinate vector in Rn, and let

W = γ|B|−1/p0∇  (xj− x0j)η x − x0 R  , (2.8)

where x0 and R are the center and radius of B, respectively. Here η is a fixed smooth function supported in B(0, 2) and identically equal to 1 on B(0, 1).

Note that since p ≤ 2, the support and bound on the L2 norm of ∇u imply

kVkLp ≤ 1, while γ can be chosen, depending only on η, n and p, so that

kWkLp0 ≤ 1. This gives us the desired properties for V and W, and moreover,

V · W = γ|B|−1/2∂u ∂xi

. (2.9)

When p ≥ 2, we need to change the definitions of V and W. As above, we take u ∈ C0∞(B) with k∇ukL2 ≤ 1, but now we set

W = |B|1/2−1/p0∇u (2.10) and (again taking j ∈ {1, . . . , n} \ {i})

V = γ0|B|−1/p ∂(ηBxj) ∂xj ei− ∂(ηBxj) ∂xi ej  , (2.11) with ηB defined by ηB(x) = η  x−x0 R 

for η as above. Clearly div V = 0 and curl W = 0, and by the conditions on u, the fact that p ≥ 2, and the choice of γ0, we can make kVkLp ≤ 1 and kWkLp0 ≤ 1. On B, where ηB is identically

equal to 1, we get V = γ0|B|−1/pe

i, so that as before

V · W = γ0|B|−1/2∂u ∂xi

. (2.12)

Integrating against g in (2.9) or (2.12), we can proceed for any p ∈ (1, ∞). Taking the supremum over all such u, we get a bound on

∂g ∂xi W−1,2(B) by a

constant multiple of |B|1/2 times the right-hand-side of (2.5). We then obtain

(7)

Case II: Now let us consider a large ball B with radius greater than 1. For this type of ball we need to show a stronger condition, that is, we need to bound the mean of |g| on B. We will do this by showing two cases of the following inequality, corresponding to (2.3) and (2.4):

 1 |B| Z B |g(x)|rdx1/r≤ C n,r sup Z g V · W. (2.13)

In both cases the supremum is to be taken over pairs of smooth vector fields V, W supported in ˜B, V ∈ (Lp)n, W ∈ (Lp0)n, with norms bounded by 1, and

in the first case, with r = p0, the vector fields satisfy condition (2.1), while in second case, with r = p, they satisfy (2.2).

Note that a priori we have the finiteness of the left-hand-side of (2.13) by the fact that g ∈ bmo ⊂ BMO ⊂ Lrloc for any r < ∞.

We first prove estimate (2.13) when B = B1is the unit ball B(0, 1). In order

to do this we need to use the full (nonhomogeneous) version of inequality (2.6), namely kgkLr(B1)≤ C ( kgkW−1,r(B1)+ n X i=1 ∂g ∂xi W−1,r(B 1) ) (2.14)

for any r, 1 < r < ∞ (see [Ne], Theorem 1, p. 108, with l = 0). The estimates for

∂g ∂xi W−1,r(B 1)

are analogous to those above (for a fixed ball the norm is equivalent to the homogeneous case). Here the test functions are in W01,r0(B1), so r0 plays the role of the exponent 2 in the argument above,

and again we need to distinguish the two cases. For r = p0, given u ∈ C0∞(B) with k∇ukLp ≤ 1, we use the same vector fields as defined in (2.7) and (2.8),

normalized for the unit ball B1. For r = p the test function u will now have

k∇ukLp0 ≤ 1, so we can define the vector fields as in (2.11) and (2.10).

Now we address the part that is different from the homogeneous case - we have to bound kgkW−1,r(B 1)= sup Z gϕ , (2.15)

where the supremum is taken over ϕ ∈ C0∞(B1), kϕkW1,r0(B1)≤ 1. We need to

be able to write any such function ϕ in terms of div-curl quantities V · W.

Lemma 2.3 If ϕ ∈ C∞

0 (B1) then we can write

ϕ = CV1· W1= CV2· W2

for smooth vector fields Vi, Wi with V1and W2supported in B1, W1 and V2

supported in the double ball fB1 = B(0, 2), W1 ∈ (Lp

0

( fB1))n, V2 ∈ (Lp( fB1))n

(with norms bounded by a constant independent of ϕ), and satisfying

div V2= 0, curl W1= 0.

(8)

(i) V1∈ (Lp(B1))n and div V1∈ Lp(B1) with bounds

kV1kLp≤ CkϕkLp, kdiv V1kLp≤ Ck∇ϕkLp;

(ii) W2∈ (Lp

0

(B1))n and, for all i, j, (curl W2)ij ∈ Lp

0

(B1) with bounds

kW2kLp0 ≤ CkϕkLp0, k(curl W2)ijkLp0 ≤ Ck∇ϕkLp0.

Proof: [Proof of Lemma 2.3:] Let ϕ ∈ C0∞(B1). Take

V1= ϕe1= (ϕ, 0, . . . , 0),

and

W1= ∇(ηx1),

where η is supported in fB1= B(0, 2), satisfies kηkL∞ ≤ 1, k∇ηkL∞ ≤ 1, and is

identically equal to 1 on the support of ϕ. Note that this ensures that on B1,

W1= e1, so we get the desired identity

V1· W1= ϕ.

We can bound the norm of W1 by a constant:

kW1kLp0(B1)≤ k∇ηk∞ Z f B1\B1 |x1|p 0 !1/p0 + kηk∞|B1|1/p 0 ≤ Cn,p,

and since W1 is a gradient, we also have curl W1= 0.

Moreover kV1kLp(B 1)≤ kϕkLp(B1) and kdiv V1kLp(B1)= ∂ϕ ∂x1 Lp(B1) ≤ k∇ϕkLp(B1).

For the other case, we can take

V2= ∂(ηx1) ∂x1 e2− ∂(ηx1) ∂x2 e1=  −x1 ∂η ∂x2 , x1 ∂η ∂x1 + η, 0, . . . , 0 

with η as above. Then

kV2kLp( fB

1). k∇ηkLp( fB1)+ kηkLp( fB1)≤ Cn,p

and div V2= 0. We also set

W2= ϕe2,

so as above, since η is identically 1 on the support of ϕ, we have

(9)

In addition,

kW2kLp0(B

1)≤ kϕkLp0(B1)

and

k(curl W2)ijkLp0(B1)≤ k∇ϕkLp0(B1).

Continuation of the proof of Theorem 2.2: We obtain estimate (2.13) for the unit ball, with r = p0 and condition (2.1) satisfied, by using (2.14), applying the lemma with the vector fields V1and W1 to the test functions in (2.15), and

dividing by the appropriate constants. The case r = p and (2.2) corresponds to using the lemma with V2 and W2.

Now let us prove estimate (2.13) for a ball B = B(x0, R), R ≥ 1. On the

left-hand-side, any g ∈ Lr(B) corresponds in a one-to-one and onto fashion to

a ˜g ∈ Lr(B 1) by taking ˜g(x) = g(x0+ Rx), with  1 |B1| Z B1 |˜g(x)|rdx 1/r = 1 |B| Z B |g(y)|rdy1/r.

On the right-hand-side, for any smooth vector fields Vi, Wi corresponding

to the unit ball, as in Lemma 2.3 (i = 1 in the case r = p0, i = 2 in the case r = p), define

V0= Vi(R−1(x − x0)), W0= Wi(R−1(x − x0)).

Then V0, W0are smooth vector fields supported in eB = B(x0, 2R), V0∈ (Lp)n,

and W0∈ (Lp 0 )n with bounds kV0kLp= Z |Vi(R−1(x − x0))|pdx 1/p = Rn/pkVikLp≤ Rn/p, and similarly kW0kLp0 ≤ Rn/p 0 . Moreover, if i = 1 we have div V0∈ Lp,

kdiv V0kLp= Z 1 R(div V1) x − x0 R  p dx 1/p = Rn/p−1kdiv V1kLp≤ Rn/p−1, and curl W0= R−1(curl W1)(R−1(x − x0)) = 0,

while if i = 2 we have div V0= 0 and, for all j, k ∈ {1, . . . , n},

k(curl W0)jkkLp= Z 1 R((curl W2)jk) x − x0 R  p0 dx 1/p0 ≤ Rn/p0−1.

(10)

Letting V = R−n/pV0, W = R−n/p

0

W0and using the fact that R ≥ 1, we

get vector fields satisfying either (2.1) in the first case, or (2.2) in the second case, and Z g V · W = R−n Z ˜ g(R−1(x − x0))Vi(R−1(x − x0)) · Wi(R−1(x − x0))dx = Z ˜ g Vi· Wi.

Taking the supremum over all such Vi, Wi, and using estimate (2.13) for the

unit ball, we get the same inequality for the ball B. This concludes the proof of the theorem.

Remarks:

1. All vector fields constructed in the proof are smooth with compact sup-port, so the suprema on the right-hand-side of (2.3) and (2.4) can be restricted to dot products of such vector fields.

2. As pointed out, the proof of Case I is independent of the radius of the ball and therefore we have generalized the result (1.4) from [CLMS] to p 6= 2. Such a generalization and the resulting decomposition of H1(Rn) in terms of (smooth) div-curl atoms is stated in [BIJZ] (Proposition 2.2) without proof.

3

The decomposition theorem for h

1

Since DC1,0p (respectively DC0,1p ) is a bounded symmetric subset of h1(Rn), we can use Lemmas III.1 and III.2 in [CLMS], the duality of bmo and h1[Go], and

Theorem 2.2 to obtain the following decomposition of functions in h1 in terms

of the appropriate “div-curl atoms”:

Theorem 3.1 For a function f in h1

(Rn), 1 < p < ∞, there exists a sequence

{λk} ∈ `1 such that f = ∞ X k=1 λkfk, fk∈ DC1,0p for all k ≥ 1.

Such a decomposition also holds with fk ∈ DC p

(11)

References

[BIJZ] Bonami, A., Iwaniec, T., Jones, P., Zinsmeister, M.,: On the prod-uct of functions in BMO and H1, Ann. Inst. Fourier (Grenoble) 57

(2007), pp. 1405–1439.

[CLMS] Coifman, R., Lions, P.-L., Meyer,Y., Semmes, S.: Compensated com-pactness and Hardy spaces, J. Math. Pures Appl. 72 (1993), pp. 247–286.

[D] Dafni, G.: Nonhomogeneous div-curl lemmas and local Hardy spaces, Adv. Differential Equations 10 (2005), pp. 505–526.

[DL] Duvaut, G., Lions, J.-L.: Inequalities in mechanics and physics, Translated from the French by C. W. John, Springer-Verlag, Berlin-New York, 1976.

[FS] Fefferman, C., Stein, E. M.: Hp spaces of several variables, Acta Math. 129 (1972), pp. 137–193.

[GR] Girault, V., Raviart, P.-A.: Finite Element Methods for Navier-Stokes Equations: Theory and Algorithms, Springer-Verlag, Berlin, 1980.

[Go] Goldberg, D.: A local version of real Hardy spaces, Duke Math. J. 46 (1979), pp. 27–42.

[Mu] F. Murat, Compacit´e par compensation, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), Vol. 5 (1978), 489–507.

[Ne] J. Neˇcas, Sur les normes ´equivalentes dans WpK(Ω) et sur la coerci-tivit´e des formes formellement positives, in “ ´Equations aux D´eriv´ees partielles”, S´emin. Math. Sup. 1965, Presse Univ. Montr´eal, Montr´eal 1966.

參考文獻

相關文件

6 《中論·觀因緣品》,《佛藏要籍選刊》第 9 冊,上海古籍出版社 1994 年版,第 1

The first row shows the eyespot with white inner ring, black middle ring, and yellow outer ring in Bicyclus anynana.. The second row provides the eyespot with black inner ring

The proof is based on Hida’s ideas in [Hid04a], where Hida provided a general strategy to study the problem of the non-vanishing of Hecke L-values modulo p via a study on the

Robinson Crusoe is an Englishman from the 1) t_______ of York in the seventeenth century, the youngest son of a merchant of German origin. This trip is financially successful,

fostering independent application of reading strategies Strategy 7: Provide opportunities for students to track, reflect on, and share their learning progress (destination). •

Now, nearly all of the current flows through wire S since it has a much lower resistance than the light bulb. The light bulb does not glow because the current flowing through it

Wang, Solving pseudomonotone variational inequalities and pseudocon- vex optimization problems using the projection neural network, IEEE Transactions on Neural Networks 17

Hope theory: A member of the positive psychology family. Lopez (Eds.), Handbook of positive