1. Introduction
In recent years, function spaces with variable exponents have been intensively studied by an important number of authors. The generalized Lebesgue spaces
(or variable exponent Lebesgue spaces) appeared in literature for the first time already in an article by Orlicz [1] , but the advanced development started with the paper [2] of Kovacik and Rakosnik in 1991. A survey of the history of the field with a bibliography of more than a hundred titles published up to 2004 can be found in [3] . To illustrate the importance of Wiener amalgams, let us mention one specific example, which today plays a central role in the theory of time-frequency analysis. This is the space
consisting of functions that are locally the Fourier transform of an
function and have a global behavior
.
The motivation to study such function spaces comes from applications to fluid dynamics [4] [5] , image processing [6] , PDE (Partial Differential Equation) and the calculus of variation [7] [8] .
In the early 1980s, in a series of articles, Feichtinger provides the most general definition of Wiener Amalgam (WA) [9] [10] [11] .
For an introduction to WA on the real line and for some historical notes, we refer to [12] .
In mathematical domain, Wiener amalgams proved to be a very useful tool, for instance in time-frequency analysis [13] (e.g. the Balian-Low theorem [12] ) and sampling theory. Our interest in those spaces arose from the Wiener Amalgams of the spaces with constant exponents [14] .
F. Holland began his systematic study in 1975 [15] . Since, he has been widely studied by [16] [17] [18] .
Only some papers treat the Wiener amalgam with one variable exponent [19] [20] [21] .
It seems that Wiener amalgams with two or more variable exponents have not yet been considered in full generality. In this work, we define a two-variable exponent amalgam space
and give some properties and study their dual.
Some properties of variable exponent amalgam space can be derived in the same way as for usual amalgams
, where
are constant, while others are very complicate.
The following definitions and results on the amalgam spaces
with constant exponents can be found in [15] [17] [22] [23] [24] [25] [26] .
1) Classical Wiener Amalgam space
with constant exponents
We give d as a fixed positive integer and
as the d-dimensional Euclidean space equipped with its Lebesgue measure dx.
For
, the amalgam of
and
is the space
defined by:
where for
(1)
with
and
.
The map
denotes the usual norm on Lebesgue space
on
while
stands for the characteristic function of the subset E of
.
2) Some basic facts about amalgam spaces
with constant exponents
Let
. Amalgam spaces
are defined in (1).
Here are the well-known results properties (see, for example, [15] [17] [22] [23] [24] [25] [26] ):
· For
,
is a norm on
equivalent to
(the equivalence constants depend only on r).
With respect to these norms, the amalgam spaces
are Banach spaces.
· The spaces are strictly increasing with the global exponent p and (strictly) decreasing with a growing local exponent q more precisely:
*
(2)
that is:
*
*
(3)
that is:
· For
, Holder’s inequality is fulfilled:
(4)
where
are conjugate exponents of
that is
.
3) Duality of the wiener amalgam spaces
with constant exponents
· When
,
is isometrically isomorphic to the dual
of
in the sense that for any element T of
, there is an unique element
of
such that:
and furthermore
.
We recall that
.
· If
, then there exist real numbers A and B such that:
for
,
and
,
.
4) Denseness of some subsets in amalgam spaces
with constant exponents
4.1) We define
to be the collection of all simple functions, that is, functions whose range is finite:
if:
where the numbers
are distinct and the sets
are pairwise disjoint.
4.2) Let Ω be an open non void set. Suppose that
, then
and
are dense in
.
5) Constant Lebesgue sequence spaces
a) For any real sequence
,
(5)
b)
(6)
c)
Therefore, we get the following proposition.
Proposition 1.
Let
.
a) Endowed with the two usual operations,
is a real vector space and the mapping
makes it a Banach space:
b) Holder’s inequality: If
and
, then
(7)
c) Suppose that
. Then the topological dual of
is isomorphically isometric to
and the duality bracket is defined as follows:
Furthermore, the following result is well-known.
Proposition 2.
a)
is a closed sub vector space of
whose topological dual is
b) For any
:
(8)
therefore
is continuously embedded in
.
Given a normed vector space V, we denote by
the normed vector space of bounded linear functionals
endowed with the usual operator norm. We wish to study
. The study is motivated by norm conjugate inequality (Theorem 22). More precisely, for
, we define the integral operator associated to g to be the operator
given by
Hölder’s inequality ensures that
is a well-defined operator and that it is bounded. The linearity of the integral implies the linearity of
whence
.
We have thus defined an operator
. Again using the linearity of the integral, we find that T is linear. What’s more, by Proposition 20, we have the identity:
From this, it follows that T is a bijection, bounded, linear operator from
into
, it actually turns out that this operator is an isomorphism.
The paper is divided into four sections. Section 2 includes fundamental notations and definitions, which will be used in the subsequent sections. Section 3 contains auxiliary results and properties. Section 4 deals with the dual of
.
Throughout the paper, the constants are independent of the main parameters involved, with values that may differ from line to line.
2. Definitions and Notations
· d will be a fixed positive integer, Ω a non void subset of
, for any subset E of
the d-dimensional euclidean space
is equipped with its Lebesgue measure
and
will be the characteristic function of E, for any
,
will be the usual euclidean norm of x.
*
in general indicate that
are functions used as norm indexes (
).
*
in general mean that
are functions which are applied on the elements
of
, the dots between the brace refer to these elements.
Let
be the set of all Lebesgue measurable functions
. In order to distinguish between variable and constant exponents, we will always denote exponent functions by
.
Given
and a set
, let:
We simply write:
As in the case for the classical Lebesgue spaces, we will encounter different behaviors depending on whether:
Therefore, we define three canonical subsets of Ω:
Below, the value of certain constants will depend on whether these sets have positive measure; if they do we will use the fact that, for instance,
Given
, we define the conjugate exponent
by:
with the convention
.
Since
is a function, the notation
can be mistaken for the derivative of
, but we will never use the symbol <<‘>> in this sense.
The notation
will always denote the conjugate of a constant exponent. The operation of taking the supremum/infimum of an exponent does not commute with forming the conjugate exponent. In fact, a straightforward computation shows that:
For simplicity, we will omit one set of parentheses and write the left-hand side of each equality as:
We will always avoid ambiguous expressions such as
.
A function
is locally log-Holder continuous and denotes this by
, if there exists a constant
, such that:
We say that
is log-Holder continuous at infinity and denote this by
, if there exist
and
such that
If
is log-Holder continuous locally and at infinity, we will denote this by writing
.
If there is no confusion about the domain we will sometimes write:
or
.
· Let
be the vector space of equivalence modulo dx-always everywhere equality of real-valued measurable functions on Ω.
· For any
and a Lebesgue measurable function f, we denote:
(9)
where
We define:
(10)
(11)
· If f is unbounded on
or
, we define
· If
in particular when
, we let
.
· If
then
.
Let I be a non void countable set,
be the set of all Lebesgue measurable functions
.
· For any
and
, we define the modular
by:
(12)
or
· If
or
is unbounded on
, we define
.
· If
, in particular when
, we let
therefore
.
· If
then
.
Definition 3.
Let I be a non-void countable set,
be the set of all functions
.
For any
, we define the variable sequence spaces
by:
(13)
where
(14)
(15)
Then, for any
,
:
(16)
We define on
some operations as follows:
For any
,
,
,
:
;
;
;
.
We also define the absolute value of any element
of
by:
the s-power of
of
(with
) is defined by:
Remark that:
· If
,
then
.
· If
,
then
.
Properties 4.
1) Let’s prove that:
(17)
Remark that (17) generalizes (8).
· Case 1:
Let’s prove that:
a.e. on I
.
If
belongs to the left-hand side set, then:
(18)
this implies that
(19)
Since
, this inequality with (19)
then
, therefore
, this implies that
then
.
· Case 2:
;
.
and
In this case:
Let’s compare
and
Take
in the first (left-hand side) set, then
, then
this implies that
belongs to the right-hand side set, therefore:
We have:
2) Given a non-void countable set I and
such that
, then for all s such that
, we have:
To prove this, let
3) When
,
, the definition (13) is equivalent to the classical norm of
seen in (6), let’s prove it:
For
,
then
and
,
For
,
then
and
, therefore
4) Given a countable and non-void set I and
, for all:
and
, if
,
then
(20)
This inequality can be generalized in the following way:
5) Given a countable and non-void set I and
, define
by:
Then, there exists a constant K such that for all:
and
(21)
In fact, we have the following result.
Proposition 5.
Given a non void countable set I and
.
a) Suppose that
. Then:
•)
is a continuous linear functional on
and
••) If
on I, then
b) Suppose that
such that
Then,
c) Suppose that
on I and T belongs to the dual
of
.
Then,
there exists
such that:
For the proof of (20) and (21) and Proposition 5, consult Theorem 2.26 and Corollary 2.28 of [27] take account of the fact that
is in fact the Lebesgue space
where
,
the power of the set X,
is defined as:
,
is a counting measure.
In this work, we will need the following lemma called Norm-modular unit ball property.
Lemma 6. (Norm-modular unit ball property)
Let Ω be a non void set
, suppose that
. For any sequence
and
,
if only if
.
Remark that the discrete version of this lemma is also valid.
Historic of the definition
Recall that: For
,
,
The amalgam of
and
is the space
defined (see (1)) by:
where
Suppose that q is a function
and p a constant and taking account of (10), we get:
(22)
is real sequence indexed by a countable set
.
Suppose that q and p are both functions
and
and taking account of (14), (22) becomes:
which may be rewritten with more information under the form:
(23)
where
Definition 7.
Let Ω be a set such that
, for any
,
, let
,
, with
,
, for any Lebesgue measurable function f we define the non negative real number:
(24)
where
If we take
in (24), we get:
(25)
We define the two-variable exponential amalgam spaces
by:
(26)
If there is no confusion:
will be smply
.
Explanation
To compute
, we first calculate:
,
(where
), this result depends at least on k and r, we denote it by
, after that we consider
and determine:
.
where
, this result depends at least on r.
Finally, the result of the calculation of
depends at least on r.
In other hand,
(27)
· A sequence
of
is said to converge in norm to f, we note:
· For two functions (eventually constants)
on Ω such that
, we define:
This is a Banach space with the norm:
3. Properties
In this section, we will use a method to show that
is a Banach space either Ω is bounded or unbounded.
Proposition 8.
Let Ω be a set such that
, given
,
and
.
1) Then
is a vector space and
2) The function
defines a norm on
.
Proof.
1)
which is a vector space, then it will suffice to show that for all
not both 0; and
:
From triangle inequality of
, we have:
is order preserving, then:
From triangle inequality of
:
From the homogeneity of
and
, we have:
that is
(28)
It is obvious that
.
2) Let
.
It is easy to see that:
, let
.
Homogeneity of
:
Let
, and
, by homogeneity of
and
, we have:
Triangle inequality of
:
In (28), if we take
, we will get:
¨
We will need the following lemmas.
Lemma 9. [28]
Let
be a measure space such that
.
Then,
that is
for any
.
Where
Lemma 10. (Monotone Convergence)
Let Ω be a set such that
, given
,
and
.
If
is a sequence of non negative functions such that
increases to a function f pointwise always everywhere (a.e.).
Then:
either
and
or
Proof.
First:
Let’s decompose f as:
(29)
where
,
.
Therefore,
We have that:
in other hand, for any non negative real numbers
:
Then,
Therefore, for any
, we can decompose it as
such that:
that is:
(30)
where
.
Now, we begin the proof:
increases to the function f a.e. (by hypothesis), we can estimate
, for any
and
:
(31)
(32)
To estimate
, we use (29) and (31), to get:
and
that is
Now, if we use Lemma 9, since
and
, we get:
But
, therefore
(33)
by the same way, we also have that:
(34)
Substituting (33) and (34) in
, we get:
Thus,
this implies that:
increases to a function f pointwise, a.e., then
for any n, since
, we have:
that is:
(35)
converges pointwise to f always everywhere i.e.
, therefore
for any
, if n is sufficiently large, (35) gives:
which implies for n sufficiently large:
, then
therefore:
as
If we replace
by its value in (32), we get:
as
.
Replacing
by its value in (31), we will find:
as
,
therefore
as
¨
Remark 11.
In the calculation of
(in the above proof), we allow the possibility
, it is the case when
or
(
) then
.
Remark 12.
If
, we have defined
, so in every case, we may write
.
Lemma 13. (Lemma of Fatou)
Let Ω be a set such that
, given
,
and
.
If
is a sequence of non negative functions such that
pointwise a.e.
If
.
Then,
and
.
Proof.
Define a sequence
,
.
Then, for all
,
and so
. By definition,
is an increasing sequence and
a.e
.
Therefore, by monotone convergence lemma:
, therefore
.
¨
Lemma 14. (Lemma of Riesz-Fischer)
Let Ω be a set such that
, given
,
and
.
If
is a sequence such that:
.
Then, there exists
such that:
in norm as
and
.
Proof.
Define the function F on Ω by:
and define the sequence
by:
.
The sequence
is non-negative and increases pointwise almost everywhere to F. Further, for each i,
and its norm is uniformly bounded, since
by hypothesis
By the monotone convergence theorem
. In particular from Proposition 8-1) F is finite a.e.
Hence, if we define the sequence
by
.
Then, this sequence also converges pointwise almost everywhere since absolute convergence implies convergence. Denote its sum by f (
as
).
Let
, then for any
,
pointwise almost everywhere.
Furthermore,
. By Fatou’s lemma, if we take
then:
More generally, for each j, the same argument shows that:
.
Since the sum in the right-hand side tends to zero, we see that
in norm, which completes the norm.¨
Proposition 15.
Let Ω be a set such that
, given
,
.
is a Banach space
Proof.
It is sufficient to show that every Cauchy sequence in
converges in norm.
Let
be a Cauchy sequence.
Choose
such that:
for
Choose
such that:
for
and so on...
This construction yields a subsequence
such that:
Define a new sequence
by:
Then, for all j, we get the sum:
Further, we have that:
Therefore, by the Riesz-Fischer lemma, there exists
such that:
in norm.
Finally, by the triangle inequality, we have that:
Since
is a Cauchy sequence, for n sufficiently large we can choose
to make the right-hand side as small as desired.
Hence,
in norm.
¨
We will need the following lemma.
Lemma 16. [27]
Given
.
1)
In particular, if Ω is bounded set:
2)
for some
.
In particular, the embedding holds if
.
Proposition 17.
Let Ω be a set such that
, given
,
and
.
1)
If
,
and
, then
2)
•) Let
.
If
on Ω, then
and
or
.
••) In particular, when
, we have:
3)
•) Let
and
on
.
Then,
and
or
.
••) In particular when
, we have:
4)
If
both in
then
with constant exponents.
with constant exponents have been widely studied by many researchers (see [1] [15] [16] [17] [18] ).
5)
If
. Then, there exist positive constant reals
such that:
otherwise,
.
6)
If
Then,
7)
Let
.
Then,
with
and
,
otherwise,
.
8)
For any
, the norms
and
are equivalent.
Proof.
1) Under the hypotheses of 1), we know that
:
Combining these two results, we will get:
, using properties 4-2), we get:
, therefore
that is
2)
•)
on Ω
on Ω, therefore, there exists
on Ω such that
on Ω, from Holder’s inequality
, but
, therefore
(Lemma 2.39 of [27] ), then we get:
,
that is
or
.
this result generalizes (3).
••) In the particular case, when
, we have:
then follows the inequality:
3)
•)
. Under the hypotheses of 3), since
on
, we have from (17):
that is
or
.
This result generalizes (2)
••)
,
if we apply 2) ••) and3) •), we get:
4)
We know that:
.
If
, then from (5) this last equality becomes:
.
Suppose that both
and
are constants belonging to
, the last equality gives:
, see (1).
We conclude that our space
generalizes both:
studied in [19] [20] [21] ; and
in [16] [17] [18] .
5)
We have
, we will use 2) and 3) to get:
or
otherwise,
6)
, from Holder’s inequality:
(36)
Case 1:
This implies that
.
is order preserving, therefore the last inequality (36) implies that:
now we apply (21) with
to get: