1. Introduction
Sequence spaces
of variable exponent
were introduced in 1931 by W. Orlicz [
1] when addressing a question related to Fourier series. In the same work, Orlicz generalized the idea of Lebesgue spaces of variable integrability by defining the class of measurable functions
f such that
Lebesgue spaces of variable exponent arise naturally in the study of hydrodynamic equations that describe the behavior of non-Newtonian fluids [
2,
3]. Electrorheological fluids, characterized by dramatic and sudden changes in viscosity when exposed to an electric or magnetic field, are typical examples. A vigorous mathematical research effort is being devoted to electrorheological fluids and their applications to civil engineering, military science and medicine, among others [
4,
5,
6,
7].
Though variable exponent Lebesgue spaces appeared for the first time in [
1], they were first studied as Banach spaces in [
8]. The core observation of our approach is the immediate effect the variability of the exponent
p on the topology of
, namely the modular structure engenders a topology that differs fundamentally from that induced by the Luxemburg norm. This phenomenon is only visible when the exponent
p is not constant (in the classical case of constant
p, the modular is topologically equivalent to the norm) and is particularly striking in the endpoint cases, namely when the exponent is finite everywhere but is either unbounded or takes up values arbitrarily close to 1.
Modular uniform convexity and its applications to fixed point theory and proximinality are now well-understood in the case when the exponent
p is unbounded, i.e., if
, [
9], but remains unexplored in the remaining endpoint situation, namely the case when
. This work aims at remedying this situation. Proposition 2 and Theorems 3 and 4 are our main results and fill the existing gap in the literature for
.
This work of exploiting the modular structure of
, new convexity properties of the modular, are discussed in the endpoint case
and concrete applications of these new properties to modular fixed point theory are presented. The reader interested in the investigation of partial differential equations in the variable exponent spaces may consult the books [
10,
11].
2. Modular Vector Spaces
Orlicz’s ideas have inspired the research of many mathematicians. In particular, his work published in 1931 [
1] inspired Nakano to introduce the concept of modular vector spaces.
Definition 1. ([
12]).
Let X be a linear vector space over the field . A modular on X is a function satisfying the following conditions:- (1)
if and only if ;
- (2)
, if ;
- (3)
, for any and any .
for any and , then ρ is called a convex modular. In addition, ρ is said to be left-continuous if for any .
It is possible to associate the idea of convergence to any modular on a vector space, in the following fashion:
Definition 2. A sequence is said to ρ-converge to if Let
,
and let
be
-convergent to
y. If the inequality
holds,
is said to satisfy the Fatou property.
A modular function on a vector space X engenders a norm in a natural fashion.
Definition 3. ([
13]).
Given a convex modular ρ defined on the vector space X, the modular space generated by ρ is the setThe functional on X, , defined by is a norm, called the Luxemburg norm.
The following major point to be underlined at this juncture is as follows: though convergence in the sense of the Luxemburg norm implies modular convergence (introduced in Definition 2), the converse fails. Deep problems arise in the study of modular spaces in which both notions of convergence are different. One concrete such instance are the Lebesgue spaces of variable, unbounded exponents, which will be considered in the present work.
3. Variable Exponent Lebesgue Spaces
Due to their applications to the hydrodynamics of smart fluids, partial differential equations with non-standard growth have been intensively studied in the past two decades (see [
2,
3,
8,
14] and the references therein). Problems of interest include the existence and uniqueness of solutions and the corresponding regularity of the solutions, when they exist. The simplest differential operator with non-standard growth is the variable exponent
p-Laplacian. Specifically, let
be open and connected with boundary
and let
stand for the class of measurable functions
. The
-Laplacian operator is defined as
Not surprisingly, the behavior of
is highly sensitive to the quality of the function
p. Existence and uniqueness results for the Dirichlet problem
are known [
15] under certain conditions for
f and
under the assumption that
p is bounded away from 1 and
∞, that is, assuming
The Restriction (
2) is standard in the literature due to the inherent limitations of standard methods in the cases
or
. In an ongoing project, two of the authors of this article showed that Theorem 2 (see below) and its implications are the right tool to handle the case of the exponent
being unbounded (but still with
) on the domain
. In fact, the authors show there that modular uniform convexity lies at the heart of the solvability of the Dirichlet problem (
1) for
as long as
.
The ultimate aim of this work is to present a theory that includes the limit case
, similar to the one developed in [
9] for
, having in mind the goal of solving the boundary value problem (
1) in the still open case
. Substantial progress is being made in this direction for the results in the present article to be viewed as the fundamental mathematical foundation of such endeavor.
Denote by
the vector space of all real-valued, Borel-measurable functions defined on an open and connected domain
with boundary
. Recall
The Lebesgue measure of a subset
will be denoted by
. For each such
p define the following sets:
Theorem 3. ([
8,
10]).
Fix in . The function defined byis a convex, continuous modular on .
The space
is defined as
endowed with the Luxemburg norm, i.e., for
,
It has been proved in [
8,
14] that
is a Banach space and that, if
p is constant on
then
coincides with the original Lebesgue space
.
The interested reader can refer to [
8,
10,
14] for a detailed treatment of these generalized Lebesgue spaces. A further note is in order at this point:
is the Musielak–Orlicz space corresponding to the Musielak–Orlicz function
given by
The Musielak–Orlicz spaces were introduced by Nakano in 1950 [
12]; the works [
13,
14] contain further information on this area of mathematics.
The modular associated to a constant exponent p is none other than the p-th power of the Luxemburg norm; it is thus clear that the obstacles found in the normed-space structure of are essentially the same as those found in its modular structure. However, if p is non-constant, the handling of the geometry of the norm presents serious technical difficulties, especially in the end-point cases for the exponent p. For this reason, the modular structure presents a viable alternative in this case.
The following lemma, of a technical nature, plays a key role in our development.
Lemma 1. The following inequalities hold:
- (i)
for any .
- (ii)
for any such that .
We refer the reader to [
9] for a detailed proof of the following result:
Theorem 2. For a bounded domain , let satisfy . Fix , and such that , and . Then, it holds the bound 4. Uniform Decrease Condition
In this section a class of subsets of
is introduced. Subsets in this class satisfy modular geometric properties similar to those of
when
. The following notations will be used:
where
.
For any measurable subset
I of
and
, define the functional
Here, it is assumed that if .
Definition 4. For , C satisfies the uniform decrease condition (or C is a set) if for each , there exists such that It is clear that the condition is inherited by the subsets of a given set and it is easy to verify that if (that is ), then has only one subset, namely . This case, however, is not interesting and it will be assumed henceforth that .
On another note, it is not hard to prove that if
, then any
satisfies the condition
. To see this, let
C be such set, fix
and set
. Then,
, from which it follows that
It transpires from the above that the condition is only interesting when and is not identically equal to 1. These two conditions will be assumed from now on.
Example 1. Set . Consider the function defined by Note that and . Consider the set It is easy to verify that C is nonempty, convex, and ρ-closed. In addition, C satisfies . This can be shown by fixing . Then, there exists for which Set . Note that if and only if . Let . We have which proves our claim that C is .
Yet another class of subsets of has to be introduced in order to proceed to the characterization of sets.
Definition 5. Let with and . Consider a non-decreasing function and set to be We remark that , since . In the following lemma, some elementary properties of are presented.
Lemma 2. In the notation of Definition 5, one has the following:
- (1)
is convex.
- (2)
is symmetric, i.e., whenever .
- (3)
The set is ρ-closed. In particular, this implies that is ρ-complete.
Proof. To prove
, let
be in
with
. Fix
and
, select
It follows that
The claim follows from the arbitrariness of and the fact that the left term on the right-hand side tends to zero as . □
Proposition 1. If , , and . Then, the following conditions are equivalent:
- (i)
C satisfies the condition .
- (ii)
There exists non-decreasing, such that .
Proof. Let
satisfy the condition
. For any
, there exists
such that
. Set
It follows easily that
is well defined and that, for each
. Pick
. It will be proved that
. To this end, note that
and that if
, then it necessarily follows that
. This yields
. On the other hand, if
,one can select
. Clearly,
and
. It follows by definition of
J, that it holds the inclusion
. Consequently, from
, for all
, one infers that
that is,
. This fact will force
. Thus, one concludes
in both cases, that is,
is non-decreasing.
It will next be shown that, for
,
, it holds that
. Notice that for all
, one has
which is a direct consequence of the inequality
. The proof will follow from the consideration of two mutually exclusive cases, namely
and
.
If
, let
; clearly
, which yields
. Consequently, one has
which in turn yields
. Assume
; then,
. It follows from (
4) that there exists
such that
. An analogous reasoning shows that
Hence, , for all , i.e., as claimed.
To show the implication
, let
. Taking
it is easy to see that
which proves that
satisfies the condition
. Clearly, any subset
also satisfies
. □
By virtue of Proposition 1, the study of
subsets of
can be reduced to the consideration of the sets of the form
. In this connection, the next result is of deep importance. Plainly, it states that subsets of the type
satisfy a well-known modular geometric property known as
(see [
18]), even in the unfavorable case when
.
Theorem 3. Let with and . Let be a non-decreasing function. For all such that , and , we have the estimate for all and .
Proof. Fix
,
,
such that
,
and
. It follows from the convexity of
that
. Fix
. On account of the properties
, one has
. Therefore,
which implies
Next, define
and observe that
. Consequently,
, for all
.
From our assumptions, we have
Using Lemma 1, we obtain
which implies
Using the convexity of the modular, we have
which implies
For the second case, assume
Our assumption on
implies
For any
, we have
, which implies
Let
, we have
Using Lemma 1, we obtain
for any
. Since
,
and
we obtain
which implies
Putting both cases together, we obtain
as claimed. □
The following lemma will lead to the main result of this work.
Lemma 3. Let with and . Consider a non-decreasing function and define , for . Then, it holds Proof. Let
. For any
, we have
which implies
. Hence,
Therefore, , that is ; this is the desired result. □
The next section is devoted to the discussion of a fixed point theorem for modular non-expansive mappings.
5. Application
As an application to Theorem 3, a modular version of a fixed point result for non-expansive mappings will be discussed. For an extensive discussion on the metric fixed point theory, the interested readers are referred to [
18]. Recall that, throughout this work, we assume that the exponent function
is not identically equal to 1.
First, a proximinality property of -closed convex subsets will be presented.
Proposition 2. Let such that . Let be non-decreasing and let be ρ-closed and convex. Then, C is proximinal. In other words, for any satisfying there exists a unique such that
Proof. It is clear that no generality is lost by assuming that
. It holds
, since
C is
-closed. Now, a sequence
will be constructed in the following way: For any
, let
be defined by the condition
. It will next be shown by contradiction that
is
-Cauchy. Assume otherwise; then, for some
one can construct a subsequence
of
such that
for any
. On account of Lemma 3,
is in
, where
for
. For fixed
, it holds
Next, observe that
with
; by virtue of Theorem 3, it follows that
where
Since
and
the convexity of
C yields
Letting
, one easily concludes that
Due to the above contradiction, it follows that is -Cauchy; in conjunction with the - completeness of , one concludes that the sequence -converges to .
On the other hand,
C is convex and
-closed: it follows from these observations that
and by virtue of Fatou property, it is concluded that
Writing , it follows . The uniqueness of the point c can be readily obtained from the strict-convexity of on , which follows from . □
The next analysis concerns the intersection property known as the property . Property was first introduced in the context of metric spaces. Specifically,
Definition 6. ([
18]).
A nonempty ρ-closed convex subset C of is said to satisfy the property if for any decreasing sequence of nonempty, ρ-closed, ρ-bounded, and convex subsets of C, we have Proposition 3. Let such that . Consider a non-decreasing function . Then, has the property .
Proof. Let
be a decreasing sequence of nonempty
-closed
-bounded convex subsets of
. For any
. For each
, the
-distance from
x to
is subject to the bound
Hence, the sequence is increasing (since is decreasing) and it is bounded above by ; let . One either has or . In the first case, it follows that for any ; this in turns yields .
If
, on account of Proposition 2, one can construct a sequence
,
, such that
with
. As in the proof of Proposition 2, it follows that
is
-Cauchy; let
be its
-limit of
. By definition of
, it follows that
, which shows that
, as claimed. On another note, Fatou property yields the following inequality:
□
Remark 1. Under the assumptions of Proposition 3, the conclusion still holds for any family of nonempty, convex, -closed subsets of C, where is an upward directed index set, as long as there exists with .
For the proof of this generalized version, let
and observe that there is no loss of generality by assuming
. For
, let
be defined by the condition
Since
is upward-directed, it can be assumed that
. This yields that
and according to Proposition 3, it follows that
. It is obvious that
is
-closed and by virtue of the last statement in the proof of Proposition 3, it is clear that
Select
such that
. It will be proved that, for any
,
. Indeed, for fixed
, if for some
,
, then it obviously holds that
. It is thus sufficient to assume that, for any
,
.
is upward-directed; hence, there exists
such that for any
and
. It can further be assumed that
for any
. It holds that
. Since for all
n,
, it follows that
. Furthermore, it holds that
Thus, , which yields the existence of a unique point such that . Given that is on , it is concluded that . In particular, it follows that for any , . It follows from the fact that , that for any , ; in turn this yields . From the arbitrariness of in , one must conclude that , so that , as claimed.
The following proposition is a further accessory to the fixed point theorem for -non-expansive mappings. The following proposition is in order:
Definition 7. A set is said to have ρ-normal structure if for any closed, convex, bounded W, , that contains more than one point, there exists such that Proposition 4. Let such that . For a non-decreasing function , has ρ-normal structure.
Proof. Let
be as in Definition 7. Since
C consists of more than one point,
. Let
,
and set
For
, set
. For fixed
, Lemma 3 yields that
and
are both in
. So far, we have
Due to the arbitrariness of
, it follows that
The latter completes the proof of Proposition 4. □
The central result of this work can now be proved.
Theorem 4. Let such that . Let C be a nonempty ρ-closed, convex, and ρ-bounded subset of and assume that C satisfies . Then, any ρ-non-expansive mapping has a fixed point.
Proof. On account of Proposition 1, since
C is
, there exists a non-decreasing function
such that
. The conclusion is trivial if
C consists of just one point; it will thus be assumed that
C contains at least two distinct points, i.e., suppose that
. Define
, which yields
Remark 1 in concert with the boundedness of
C and Zorn’s lemma rapidly yields a minimal element of
, say
. In what follows, it will be shown that
consists of exactly one point. This follows by contradiction. If
consisted of more than one point, one could set
to be the intersection of all
closed convex subset of
C containing
. Clearly,
since
. Furthermore,
which, in turn, yields
. Since
is a minimal element of
it is easy to see that
. According to Proposition 4, there exists
such that
Define the subset
. It is obvious that
, also,
. A straightforward reasoning using modular balls shows that
is
closed and convex. In fact, it follows that
. Indeed, let
. Since
T is
non-expansive, for all
it holds
Thus,
, which yields
. The equality
, implies that
, which in turn yields
for all
. Thus,
, but
x is an arbitrary element of
K, so that in all,
. The minimality of
forces
. In all,
This is a contradiction. It follows that consists of only one point, which must be a fixed point of T because . □
Remark 2. The condition in Theorem 4 can be replaced with the slightly more general condition: