• Nie Znaleziono Wyników

C O L L O Q U I U M M A T H E M A T I C U M VOL. 82 1999 NO. 2

N/A
N/A
Protected

Academic year: 2021

Share "C O L L O Q U I U M M A T H E M A T I C U M VOL. 82 1999 NO. 2"

Copied!
12
0
0

Pełen tekst

(1)

C O L L O Q U I U M M A T H E M A T I C U M

VOL. 82 1999 NO. 2

FEJ ´ ER MEANS OF TWO-DIMENSIONAL FOURIER TRANSFORMS ON H

p

(R × R)

BY

FERENC W E I S Z (BUDAPEST)

Abstract. The two-dimensional classical Hardy spaces Hp(R×R) are introduced and it is shown that the maximal operator of the Fej´er means of a tempered distribution is bounded from Hp(R×R) to Lp(R2) (1/2 < p ≤ ∞) and is of weak type (H1(R×R), L1(R2)) where the Hardy space H1(R × R) is defined by the hybrid maximal function. As a con- sequence we deduce that the Fej´er means of a function f ∈ H1(R × R) ⊃ L log L(R2) converge to f a.e. Moreover, we prove that the Fej´er means are uniformly bounded on Hp(R×R) whenever 1/2 < p < ∞. Thus, in case f ∈ Hp(R×R), the Fej´er means converge to f in Hp(R × R) norm (1/2 < p < ∞). The same results are proved for the conjugate Fej´er means.

1. Introduction. The Hardy–Lorentz spaces H

p,q

(R × R) of tempered distributions are endowed with the L

p,q

(R

2

) Lorentz norms of the non- tangential maximal function. Clearly, H

p

(R × R) = H

p,p

(R × R) are the usual Hardy spaces (0 < p ≤ ∞).

In Zygmund [22] (Vol. II, p. 246) it is shown that the Fej´er means σ

T

f of a one-dimensional function f ∈ L

1

(R) converge to f a.e. as T → ∞.

Moreover, the maximal operator of the Fej´er means, σ

:= sup

T >0

T

|, is of weak type (1, 1), i.e.

sup

γ>0

γλ(σ

f > γ) ≤ Ckf k

1

(f ∈ L

1

(R))

(see Zygmund [22], Vol. I, p. 154 and M´oricz [14]). M´oricz [14] also verified that σ

is bounded from H

1

(R) to L

1

(R). The author [19] proved that σ

is also bounded from H

p,q

(R) to L

p,q

(R) whenever 1/2 < p < ∞, 0 < q ≤ ∞.

In [16] we investigated the Fej´er means of two-parameter Fourier series and proved that σ

:= sup

n,m∈N

n,m

| is bounded from H

p,q

(T × T) to

1991 Mathematics Subject Classification: Primary 42B08; Secondary 42B30.

Key words and phrases: Hardy spaces, p-atom, atomic decomposition, interpolation, Fej´er means.

This research was made while the author was visiting the Humboldt University in Berlin supported by the Alexander von Humboldt Foundation.

[155]

(2)

L

p,q

(T

2

) (3/4 < p ≤ ∞, 0 < q ≤ ∞) and is of weak type (H

1

(T × T), L

1

(T

2

)), i.e.

sup

γ>0

γλ(σ

f | > γ) ≤ Ckf k

H

1(T×T)

(f ∈ H

1

(T × T)).

Moreover, the Fej´er means σ

n,m

f converge to f a.e. as n, m → ∞ whenever f ∈ H

1

(T × T) ⊃ L log L(T

2

) (see Weisz [15], [16] and Zygmund [22] for L log L(T

2

)).

In this paper we sharpen and generalize these results for the Fej´er means of two-dimensional Fourier transforms.

We show that the maximal operator σ

is bounded from H

p,q

(R × R) to L

p,q

(R

2

) whenever 1/2 < p < ∞, 0<q ≤∞, and is of weak type (H

1

(R × R), L

1

(R

2

)). We introduce the conjugate distributions e f

(i,j)

, the conjugate Fej´er means e σ

(i,j)T,U

and the conjugate maximal operators e σ

(i,j)

(i, j = 0, 1). We prove that the operator e σ

(i,j)

is also of type (H

p,q

(R × R), L

p,q

(R

2

)) (1/2 <

p < ∞, 0 < q ≤ ∞) and of weak type (H

1

(R × R), L

1

(R

2

)).

A usual density argument then implies that the Fej´er means σ

T,U

f con- verge to f a.e. and the conjugate Fej´er means e σ

T,U(i,j)

f converge to e f

(i,j)

(i, j = 0, 1) a.e. as T, U → ∞ provided that f ∈ H

1

(R × R). Note that e f

(i,j)

is not necessarily in H

1

(R × R) whenever f is.

We also prove that the operators σ

T,U

and e σ

(i,j)T,U

(T, U ∈ R) are uniformly bounded from H

p,q

(R×R) to H

p,q

(R×R) if 1/2 < p < ∞, 0 < q ≤ ∞. From this it follows that σ

T,U

f → f and e σ

(i,j)T,U

f → e f (i, j = 0, 1) in H

p,q

(R × R) norm as T, U → ∞ whenever f ∈ H

p,q

(R×R) and 1/2 < p < ∞, 0 < q ≤ ∞.

2. Hardy spaces and conjugate functions. Let R denote the real numbers, R

+

the positive real numbers and let λ be the 2-dimensional Lebesgue measure. We also use the notation |I| for the Lebesgue measure of the set I. We briefly write L

p

for the real L

p

(R

2

, λ) space; the norm (or quasinorm) in this space is defined by kf k

p

:= (

T

R2

|f |

p

dλ)

1/p

(0 < p ≤ ∞).

The distribution function of a Lebesgue-measurable function f is defined by

λ({|f | > ̺}) := λ({x : |f (x)| > ̺}) (̺ ≥ 0).

The weak L

p

space L

p

(0 < p < ∞) consists of all measurable functions f for which

kf k

Lp

:= sup

̺>0

̺λ({|f | > ̺})

1/p

< ∞

and we set L

= L

.

(3)

The spaces L

p

are special cases of the more general, Lorentz spaces L

p,q

. In their definition another concept is used. For a measurable function f the non-increasing rearrangement is defined by

f (t) := inf{̺ : λ({|f | > ̺}) ≤ t}. e

The Lorentz space L

p,q

is defined as follows: for 0 < p < ∞, 0 < q < ∞, kf k

p,q

:=



\

0

f (t) e

q

t

q/p

dt t



1/q

, while for 0 < p ≤ ∞,

kf k

p,∞

:= sup

t>0

t

1/p

f (t). e Let

L

p,q

:= L

p,q

(R

2

, λ) := {f : kf k

p,q

< ∞}.

One can show the following equalities:

L

p,p

= L

p

, L

p,∞

= L

p

(0 < p ≤ ∞) (see e.g. Bennett–Sharpley [1] or Bergh–L¨ofstr¨ om [2]).

Let f be a tempered distribution on C

(R

2

) (briefly f ∈ S

(R

2

) = S

).

The Fourier transform of f is denoted by b f . In the special case when f is an integrable function,

f (t, u) = b 1 2π

\

R

\

R

f (x, y)e

−ıtx

e

−ıuy

dx dy (t, u ∈ R) where ı = √

−1.

For f ∈ S

and t, u > 0 let

F (x, y; t, u) := (f ∗ P

t

× P

u

)(x, y) where ∗ denotes convolution and

P

t

(x) := ct

t

2

+ x

2

(x ∈ R) is the Poisson kernel.

For α > 0 let

Γ

α

:= {(x, t) : |x| < αt},

a cone with vertex at the origin. We denote by Γ

α

(x) (x ∈ R) the translate of Γ

α

with vertex at x. The non-tangential maximal function is defined by

F

α,β

(x, y) := sup

(x,t)∈Γα(x), (y,u)∈Γβ(y)

|F (x

, y

; t, u)| (α, β > 0).

For 0 < p, q ≤ ∞ the Hardy–Lorentz space H

p,q

(R × R) = H

p,q

consists of all tempered distributions f for which F

α,β

∈ L

p,q

; we set

kf k

Hp,q

:= kF

1,1

k

p,q

.

(4)

For 0 < p < ∞, 0 < q ≤ ∞ Chang and Fefferman [3] and Lin [12] proved the equivalence kF

α,β

k

p,q

∼ kF

1,1

k

p,q

(α, β > 0). It is known that if f ∈ H

p

(0 < p < ∞) then f (x, y) = lim

t,u→0

F (x, y; t, u) in the sense of distributions (see Gundy–Stein [11], Chang–Fefferman [3]).

Let us introduce the hybrid Hardy spaces. For f ∈ L

1

and t > 0 let G(x, y; t) := 1

√ 2π

\

R

f (v, y)P

t

(x − v) dv and

G

+α

(x, y) := sup

(x,t)∈Γα(x)

|G(x

, y; t)| (0 < α < 1).

We say that f ∈ L

1

is in the hybrid Hardy–Lorentz space H

p,q

(R×R) = H

p,q

if

kf k

Hp,q

:= kG

+1/2

k

p,q

< ∞.

The equivalences kG

+α

k

p,q

∼ kG

+1

k

p,q

(α > 0, 0 < p < ∞, 0 < q ≤ ∞) and H

p,q

∼ H

p,q

∼ L

p,q

(1 < p < ∞, 0 < q ≤ ∞)

were proved in Fefferman–Stein [7], Gundy–Stein [11] and Lin [12]. Note that for p = q the usual definitions of the Hardy spaces H

p,p

= H

p

and H

p,p

= H

p

are obtained.

The following interpolation result concerning Hardy–Lorentz spaces will be used several times in this paper (see Lin [12] and also Weisz [17]).

Theorem A. If a sublinear (resp. linear ) operator V is bounded from H

p0

to L

p0

(resp. to H

p0

) and from L

p1

to L

p1

(p

0

≤ 1 < p

1

< ∞) then it is also bounded from H

p,q

to L

p,q

(resp. to H

p,q

) if p

0

< p < p

1

and 0 < q ≤ ∞.

In this paper the constants C are absolute, while C

p

(resp. C

p,q

) depend only on p (resp. p and q) and may be different in different contexts.

One can prove similarly to the discrete case (see Weisz [16]) that L log L := L log L(R

2

) ⊂ H

1

⊂ H

1,∞

, more exactly,

(1) kf k

H1,∞

= sup

̺>0

̺λ(F

1,1

> ̺) ≤ Ckf k

H1

(f ∈ H

1

) and

kf k

H1

≤ C + Ck|f | log

+

|f |k

1

(f ∈ L log L) where log

+

u = 1

{u>1}

log u.

For a tempered distribution f ∈ H

p

(0 < p < ∞) the Hilbert transforms or conjugate distributions e f

(1,0)

, e f

(0,1)

and e f

(1,1)

are defined by

( e f

(1,0)

)

(t, u) := (−ı sign t) b f (t, u) (t, u ∈ R)

(conjugate with respect to the first variable),

(5)

( e f

(0,1)

)

(t, u) := (−ı sign u) b f (t, u) (t, u ∈ R) (conjugate with respect to the second variable) and

( e f

(1,1)

)

(t, u) := (− sign(tu)) b f (t, u) (t, u ∈ R)

(conjugate with respect to both variables). We use the notation e f

(0,0)

:= f . Gundy and Stein [10], [11] verified that if f ∈ H

p

(0 < p < ∞) then all conjugate distributions are also in H

p

and

(2) kf k

Hp

= k e f

(i,j)

k

Hp

(i, j = 0, 1).

Furthermore (see also Chang and Fefferman [3], Frazier [9], Duren [5]), (3) kf k

Hp

∼ kf k

p

+ k e f

(1,0)

k

p

+ k e f

(0,1)

k

p

+ k e f

(1,1)

k

p

.

As is well known, if f is an integrable function then f e

(1,0)

(x, y) = p.v. 1

π

\

R

f (x − t, y)

t dt := lim

ε→0

1 π

\

ε<|t|

f (x − t, y)

t dt,

f e

(0,1)

(x, y) = p.v. 1 π

\

R

f (x, y − u)

u du,

f e

(1,1)

(x, y) = p.v. 1 π

2

\

R

\

R

f (x − t, y − u)

tu dt du.

Moreover, the conjugate functions e f

(1,0)

, e f

(0,1)

and e f

(1,1)

exist almost everywhere, but they are not integrable in general. Similarly, if f ∈ H

1

then e f

(0,1)

and e f

(1,1)

are not necessarily in H

1

.

3. Fej´ er means. Suppose first that f ∈ L

p

for some 1 ≤ p ≤ 2. It is known that under certain conditions

f (x, y) = 1 2π

\

R

\

R

f (t, u)e b

ıxt

e

ıyu

dt du (x, y ∈ R).

This motivates the definition of the Dirichlet integral s

t,u

f : s

t,u

f (x, y) := 1

t

\

−t u

\

−u

f (v, w)e b

ıxv

e

ıyw

dv dw (t, u > 0).

The conjugate Dirichlet integrals are introduced by es

(1,0)t,u

f (x, y) := 1

t

\

−t u

\

−u

(−ı sign v) b f (v, w)e

ıxv

e

ıyw

dv dw (t, u > 0),

es

(0,1)t,u

f (x, y) := 1 2π

t

\

−t u\

−u

(−ı sign w) b f (v, w)e

ıxv

e

ıyw

dv dw (t, u > 0)

(6)

and

es

(1,1)t,u

f (x, y) := 1 2π

t

\

−t u

\

−u

(− sign(vw)) b f (v, w)e

ıxv

e

ıyw

dv dw (t, u > 0).

The Fej´er and conjugate Fej´er means are defined by e

σ

(i,j)T,U

f (x, y) := 1 T U

T

\

0 U

\

0

es

(i,j)t,u

f (x, y) dt du (T, U > 0; i, j = 0, 1).

We write s

t,u

f =: es

t,u(0,0)

f and σ

T,U

f := e σ

T,U(0,0)

f . It is easy to see that s

t,u

f (x, y) :=

\

R

\

R

f (x − v, y − w) sin tv

πv · sin uw πw dv dw and

σ

T,U

f (x, y) :=

\

R

\

R

f (x − t, y − u)K

T

(t)K

U

(u) dt du where

K

T

(t) := 2

π · sin

2

(T t/2) T t

2

is the Fej´er kernel. Note that

(4)

\

R

K

T

(t) dt = 1 (T > 0) (see Zygmund [22], Vol. II, pp. 250–251).

We extend the definition of the Fej´er means and conjugate Fej´er means to tempered distributions as follows:

e

σ

T,U(i,j)

f := e f

(i,j)

∗ (K

T

× K

U

) (T, U > 0; i, j = 0, 1).

One can show that e σ

(i,j)T,U

f is well defined for all tempered distributions f ∈ H

p

(0 < p ≤ ∞) and for all functions f ∈ L

p

(1 ≤ p ≤ ∞) (cf.

Fefferman–Stein [7]).

The maximal and maximal conjugate Fej´er operators are defined by e

σ

(i,j)

f := sup

T,U >0

|e σ

(i,j)T,U

f | (i, j = 0, 1).

We again write σ

f := e σ

(0,0)

f .

4. The boundedness of the maximal Fej´ er operator. A function a ∈ L

2

is called a rectangle p-atom if there exists a rectangle R ⊂ R

2

such that

(i) supp a ⊂ R,

(ii) kak

2

≤ |R|

1/2−1/p

,

(7)

(iii) for all x, y ∈ R and all N ≤ [2/p − 3/2],

\

R

a(x, y)x

N

dx =

\

R

a(x, y)y

N

dy = 0.

If I is an interval then let rI be the interval with the same center as I and with length r|I| (r ∈ N). For a rectangle R = I × J let rR = rI × rJ.

An operator V which maps the set of tempered distributions into the collection of measurable functions will be called p-quasi-local if there exist a constant C

p

> 0 and η > 0 such that for every rectangle p-atom a supported on the rectangle R and for every r ≥ 2 one has

\

R2\2rR

|T a|

p

dλ ≤ C

p

2

−ηr

.

Although H

p

cannot be decomposed into rectangle p-atoms, in the next theorem it is enough to take such atoms (see Weisz [16], Fefferman [8]).

Theorem B. Suppose that the operator V is sublinear and p-quasi-local for some 0 < p ≤ 1. If V is bounded from L

2

to L

2

then

kV f k

p

≤ C

p

kf k

Hp

(f ∈ H

p

).

Since the Fej´er kernel is positive, we can prove the following inequality in the same way as in the discrete case (see Weisz [18]):

(5) kσ

f k

p

≤ C

p

kf k

p

(1 < p ≤ ∞).

Now we can formulate our main result.

Theorem 1. We have

(6) kσ

f k

p,q

≤ C

p,q

kf k

Hp,q

(f ∈ H

p,q

)

for every 1/2 < p < ∞ and 0 < q ≤ ∞. In particular, if f ∈ H

1

then

(7) λ(σ

f > ̺) ≤ C

̺ kf k

H

1

(̺ > 0).

P r o o f. First we will show that the operator σ

is p-quasi-local for each 1/2 < p ≤ 1. To this end let a be an arbitrary rectangle p-atom with support R = I × J and

2

K−1

< |I| ≤ 2

K

, 2

L−1

< |J| ≤ 2

L

(K, L ∈ Z).

We can suppose that the center of R is zero. In this case [−2

K−2

, 2

K−2

] ⊂ I ⊂ [−2

K−1

, 2

K−1

] and

[−2

L−2

, 2

L−2

] ⊂ J ⊂ [−2

L−1

, 2

L−1

].

(8)

To prove the p-quasi-locality of the operator σ

we have to integrate |σ

a|

p

over

R

2

\ 2

r

R = (R \ 2

r

I) × J ∪ (R \ 2

r

I) × (R \ J)

∪ I × (R \ 2

r

J) ∪ (R \ I) × (R \ 2

r

J) where r ≥ 2 is an arbitrary integer.

First we integrate over (R \ 2

r

I) × J. Obviously,

\

R\2rI

\

J

a(x, y)|

p

dx dy ≤ X

|i|=2r−2

(i+1)2K

\

i2K

\

J

a(x, y)|

p

dx dy.

For x, y ∈ R let A

1,0

(x, y) :=

x

\

−∞

a(t, y) dt, A

0,1

(x, y) :=

y

\

−∞

a(x, u) du and

A

1,1

(x, y) :=

x

\

−∞

y

\

−∞

a(t, y) dt du.

By (iii) of the definition of the rectangle atom we can show that supp A

k,l

⊂ R and A

k,l

is zero at the vertices of R (k, l = 0, 1). Moreover, using (ii) we can compute that

(8) kA

k,l

k

2

≤ |I|

k

|J|

l

(|I| · |J|)

1/2−1/p

(k, l = 0, 1).

Integrating by parts we can see that

T,U

a(x, y)| =

\

I

\

J

A

1,0

(t, u)K

T

(x − t)K

U

(y − u) dt du

\

I

\

J

A

1,0

(t, u)K

U

(y − u) du |K

T

(x − t)| dt.

Using the inequality

|K

T

(t)| ≤ C/t

2

(T ∈ R

+

) we get

T,U

a(x, y)| ≤

\

I

\

J

A

1,0

(t, u)K

U

(y − u) du C

|x − t|

2

dt

≤ C2

−2K

i

2

\

I

\

J

A

1,0

(t, u)K

U

(y − u) du dt

for x ∈ [i2

K

, (i + 1)2

K

). H¨older’s inequality, the one-dimensional version of

(5) and (8) imply

(9)

\

J

a(x, y)|

p

dy

≤ C

p

2

−2Kp

i

2p

|J|

1−p



\

I

\

J

sup

U ∈R+

\

J

A

1,0

(t, u)K

U

(y − u) du dy dt 

p

≤ C

p

2

−2Kp

|J|

1−p/2

i

2p



\

I



\

R

sup

U ∈R+

\

J

A

1,0

(t, u)K

U

(y − u) du

2

dy 

1/2

dt 

p

≤ C

p

2

−2Kp

|J|

1−p/2

i

2p



\

I



\

J

|A

1,0

(t, y)|

2

dy 

1/2

dt 

p

≤ C

p

2

−2Kp

|I|

p/2

|J|

1−p/2

i

2p



\

I

\

J

|A

1,0

(t, y)|

2

dy dt 

p/2

≤ C

p

2

−2Kp

|I|

2p−1

i

2p

.

Hence

\

R\2rI

\

J

a(x, y)|

p

dx dy ≤ C

p

X

i=2r−2

2

K

2

−K

i

2p

≤ C

p

2

−r(2p−1)

. Next we integrate over (R \ 2

r

I) × (R \ J):

\

R\2rI

\

R\J

a(x, y)|

p

dx dy ≤ X

|i|=2r−2

X

|j|=1 (i+1)2K

\

i2K

(j+1)2L

\

j2L

a(x, y)|

p

dx dy.

Integrating by parts we obtain, for x ∈[i2

K

, (i+1)2

K

) and y ∈[i2

L

, (i+1)2

L

),

T,U

a(x, y)| =

\

I

\

J

A

1,1

(t, u)K

T

(x − t)K

U

(y − u) dt du

≤ C2

−2K

2

−2L

i

2

j

2

\

I

\

J

|A

1,1

(t, u)| dt du

≤ C2

−2K

2

−2L

|I|

2−1/p

|J|

2−1/p

i

2

j

2

.

Thus

\

R\2rI

\

R\J

a(x, y)|

p

dx dy ≤ C

p

X

|i|=2r−2

X

|j|=1

2

K+L

2

−K

2

−L

i

2p

j

2p

≤ C

p

2

−r(2p−1)

.

The integrations over I × (R \ 2

r

J) and over (R \ I) × (R \ 2

r

J) are

similar. Hence σ

is p-quasi-local. Theorem B implies (6) for p = q. Applying

Theorem A and (5) we obtain (6).

(10)

Let us single out this result for p = 1 and q = ∞. If f ∈ H

1

then (1) implies

f k

1,∞

= sup

̺>0

γλ(σ

f > ̺) ≤ Ckf k

H1,∞

≤ Ckf k

H1

, which shows (7). The proof of the theorem is complete.

Note that Theorem 1 was proved for Fourier series and for 3/4 < p < ∞ by the author [16] with another method.

We can state the same for the maximal conjugate Fej´er operator.

Theorem 2. For i, j = 0, 1 we have

ke σ

(i,j)

f k

p,q

≤ C

p,q

kf k

Hp,q

(f ∈ H

p,q

)

for every 1/2 < p < ∞ and 0 < q ≤ ∞. In particular, if f ∈ H

1

then λ(e σ

(i,j)

f > ̺) ≤ C

̺ kf k

H1

(̺ > 0).

P r o o f. By Theorem 1 for p = q and (2) we obtain

ke σ

(i,j)

f k

p

= kσ

f e

(i,j)

k

p

≤ C

p

k e f

(i,j)

k

Hp

= C

p

kf k

Hp

(f ∈ H

p

) for every 1/2 < p < ∞. Now Theorem 2 follows from Theorem A and (1).

Since the set of those functions f ∈ L

1

whose Fourier transform has a compact support is dense in H

1

(see Wiener [20]), the weak type inequalities of Theorems 1 and 2 and the usual density argument (see Marcinkiewicz–

Zygmund [13]) imply

Corollary 1. If f ∈ H

1

(⊃ L log L) and i, j = 0, 1 then e

σ

(i,j)T,U

f → e f

(i,j)

a .e. as T, U → ∞.

Note that e f

(i,j)

is not necessarily in H

1

whenever f is.

Now we consider the norm convergence of σ

T,U

f . It follows from (5) that σ

T,U

f → f in L

p

norm as T, U → ∞ if f ∈ L

p

(1 < p < ∞). We are going to generalize this result.

Theorem 3. Assume that T, U ∈ R

+

and i, j = 0, 1. Then ke σ

T,U(i,j)

f k

Hp,q

≤ C

p,q

kf k

Hp,q

(f ∈ H

p,q

) for every 1/2 < p < ∞ and 0 < q ≤ ∞.

P r o o f. Since (σ

T,U

f )

∼(i,j)

= e σ

(i,j)T,U

f , by Theorem 2 we have k(σ

T,U

f )

∼(i,j)

k

p

≤ C

p

kf k

Hp

(f ∈ H

p

) for all T, U ∈ R

+

and i, j = 0, 1. (3) implies that

T,U

f k

Hp

≤ C

p

kf k

Hp

(f ∈ H

p

; T, U ∈ R

+

).

(11)

Hence, for i, j = 0, 1,

ke σ

(i,j)T,U

f k

Hp

≤ C

p

kf k

Hp

(f ∈ H

p

; T, U ∈ R

+

).

which together with Theorem A implies Theorem 3.

Corollary 2. Suppose that 1/2 < p < ∞, 0 < q ≤ ∞ and i, j = 0, 1. If f ∈ H

p,q

then

e

σ

T,U(i,j)

f → e f

(i,j)

in H

p,q

norm as T, U → ∞.

We suspect that Theorems 1, 2 and 3 are not true for p ≤ 1/2 though we could not find any counterexample.

REFERENCES

[1] C. B e n n e t t and R. S h a r p l e y, Interpolation of Operators, Pure Appl. Math. 129, Academic Press, New York, 1988.

[2] J. B e r g h and J. L ¨o f s t r ¨o m, Interpolation Spaces. An Introduction, Springer, Berlin, 1976.

[3] S.-Y. A. C h a n g and R. F e f f e r m a n, Some recent developments in Fourier analysis andHp-theory on product domains, Bull. Amer. Math. Soc. 12 (1985), 1–43.

[4] R. R. C o i f m a n and G. W e i s s, Extensions of Hardy spaces and their use in anal- ysis, Bull. Amer. Math. Soc. 83 (1977), 569–645.

[5] P. D u r e n, Theory of Hp Spaces, Academic Press, New York, 1970.

[6] R. E. E d w a r d s, Fourier Series. A Modern Introduction, Vol. 2, Springer, Berlin, 1982.

[7] C. F e f f e r m a n and E. M. S t e i n, Hp spaces of several variables, Acta Math. 129 (1972), 137–194.

[8] R. F e f f e r m a n, Calder´on–Zygmund theory for product domains: Hp spaces, Proc.

Nat. Acad. Sci. U.S.A. 83 (1986), 840–843.

[9] A. P. F r a z i e r, The dual space of Hp of the polydisc for0 < p < 1, Duke Math. J.

39 (1972), 369–379.

[10] R. F. G u n d y, Maximal function characterization of Hp for the bidisc, in: Lecture Notes in Math. 781, Springer, Berlin, 1982, 51–58.

[11] R. F. G u n d y and E. M. S t e i n, Hp theory for the poly-disc, Proc. Nat. Acad. Sci.

U.S.A. 76 (1979), 1026–1029.

[12] K.-C. L i n, Interpolation between Hardy spaces on the bidisc, Studia Math. 84 (1986), 89–96.

[13] J. M a r c i n k i e w i c z and A. Z y g m u n d, On the summability of double Fourier se- ries, Fund. Math. 32 (1939), 122–132.

[14] F. M ´o r i c z, The maximal Fej´er operator for Fourier transforms of functions in Hardy spaces, Acta Sci. Math. (Szeged) 62 (1996), 537–555.

[15] F. W e i s z, Ces`aro summability of one- and two-dimensional trigonometric-Fourier series, Colloq. Math. 74 (1997), 123–133.

[16] — Ces`aro summability of two-parameter trigonometric-Fourier series, J. Approx.

Theory 90 (1997), 30–45.

[17] — Martingale Hardy Spaces and Their Applications in Fourier-Analysis, Lecture Notes in Math. 1568, Springer, Berlin, 1994.

(12)

[18] F. W e i s z, Strong summability of two-dimensional trigonometric-Fourier series, Ann. Univ. Sci. Budapest Sect. Comput. 16 (1996), 391–406.

[19] —, The maximal Fej´er operator of Fourier transforms, Acta Sci. Math. (Szeged) 64 (1998), 515–525.

[20] N. W i e n e r, The Fourier Integral and Certain of its Applications, Dover, New York, 1959.

[21] J. M. W i l s o n, On the atomic decomposition for Hardy spaces, Pacific J. Math. 116 (1985), 201–207.

[22] A. Z y g m u n d, Trigonometric Series, Cambridge Univ. Press, London, 1959.

Department of Numerical Analysis E¨otv¨os L. University

P´azm´any P. s´et´any 1/D H-1117 Budapest, Hungary E-mail: weisz@ludens.elte.hu

Received 17 June 1998

Cytaty

Powiązane dokumenty

For a tubular Σ we are going to show that the group of additive functions for Σ extending to Σ with value zero at the extension vertex has rank one, and then by Proposition

third way of viewing the spaces Cen(S); this way, compact c-sets, is pre- sented in the fashion of the weakly dyadic spaces (a generalization of centered spaces) introduced by Kulpa

Further, our method of getting lower estimates (Theorem 2) requires countable boundary multiplicity since it is based on the application of the result by Belna, Colwell and

Key words and phrases: multiple solutions, discontinuous function, elliptic inclusion, first eigenvalue, p-Laplacian, Rayleigh quotient, nonsmooth Palais–Smale condition, coer-

This paper is concerned with a second-order functional differential equa- tion of the form x ′′ (z) = x(az + bx ′ (z)) with the distinctive feature that the argument of the

For simplicity, let us introduce some notation.. We also denote by

We discuss k-rotundity, weak k-rotundity, C-k-rotundity, weak C-k-rotun- dity, k-nearly uniform convexity, k-β property, C-I property, C-II property, C-III property and nearly

In this note a traditional sliding hump argument is used to establish a simple mapping property of c 0 which simultaneously yields extensions of the preceding theorems as