1. Introduction
An immense branch of functional analysis is devoted to the topological and geometric properties of topological vector spaces (see, for example, [
1,
2,
3,
4]). Studies of bases in Banach spaces compose a large part of it (see, for example, [
1,
5,
6,
7,
8,
9,
10,
11,
12] and the references therein). It is not surprising that for concrete classes of Banach spaces, many open problems remain, particularly for the Müntz spaces
, where
(see [
13,
14,
15,
16,
17,
18,
19,
20,
21,
22] and the references therein). These spaces are defined as completions of the linear span over
or
of monomials
with
on the segment
relative to the
norm, where
,
. In his classical work, K.Weierstrass had proven in 1885 the theorem about polynomial approximations of continuous functions on the segment. However, the space of continuous functions also forms an algebra. Generalizations of such spaces were considered by C. Müntz in 1914, such that his spaces did not have the algebra structure. C. Müntz considered conditions on the exponents
for which the monomials
span a dense subspace of
. Naturally, a problem arose whether they have bases [
23,
24]. Then, the progress was for lacunary Müntz spaces satisfying the condition
with a countable set Λ, but in its generality, this problem was not solved [
20]. It is worth mentioning that the system
itself does not contain a Schauder basis for a non-lacunary set Λ satisfying the Müntz and gap conditions.
In
Section 2, relations between Müntz spaces satisfying the Müntz and gap conditions are considered. A Fourier approximation of functions in the Müntz spaces
of
functions is studied in
Section 3, where
. Necessary definitions are recalled. It is proven that up to an isomorphism and a change of variables, these spaces are contained in Weil–Nagy’s class. For this purpose, in Lemmas 1 and 2, Theorem 1 and Corollary 1, some isomorphisms of Müntz spaces are given. Then, in Theorem 2, a relation between Müntz spaces and Weil–Nagy’s classes is established. Moreover, in
Section 4, the existence of Schauder bases in the Müntz spaces
is investigated (see Theorem 3) with the help of Fourier series approximation (see Lemma 5). It is proven that, under the Müntz condition and the gap condition, Schauder bases exist in the Müntz spaces
, where
.
All main results of this paper are obtained for the first time. They can be used for further investigations of function approximations and the geometry of Banach spaces. It is important not only for the development of mathematical analysis and of functional analysis, but also in their many-sided applications.
2. Relation between Spaces
To avoid misunderstandings, we first remind about the necessary definitions and notations.
Notation 1. Let denote the Banach space of all continuous functions supplied with the absolute maximum norm:where , is either the real field or the complex field .
Then,
denotes the Banach space of all Lebesgue measurable functions
possessing the finite norm:
where
is a marked number,
.
As usual, will stand for the linear span of vectors over a field .
Definition 1. Take a countable infinite subset in the set so that is a strictly increasing sequence.
Henceforth, it is supposed that the set Λ satisfies the gap condition:
- (1)
and the Müntz condition:
- (2)
The completion of the linear space containing all monomials with and and relative to the norm is denoted by , where , , also by when it is completed relative to the norm. Briefly, they will also be written as or , respectively, for and , when is specified.
Before the subsections about the Fourier approximation in Müntz spaces auxiliary, Lemmas 1 and 2 and Theorem 1 are proven about isomorphisms of Müntz spaces . With their help, our consideration reduces to a subclass of Müntz spaces so that a set Λ is contained in the set of natural numbers .
Lemma 1. For each , the Müntz spaces and are linearly topologically isomorphic, where .
Proof. For every
and
and
, the norms
and
are equivalent, where
for
. Due to the Remez-type and the Nikolski-type inequalities (see Theorem 6.2.2 in [
16] and Theorem 7.4 in [
17]) for each Λ satisfying the Müntz condition, there is a constant
, so that
for each
, where
is independent of
h. Therefore, the norms
and
are equivalent on
. Certainly, each polynomial
defined on the segment
has the natural extension on
, where
are constants and
t is a variable. Thus, the Müntz spaces
and
are linearly topologically isomorphic as normed spaces for each
. ☐
Lemma 2. The Müntz spaces and are linearly topologically isomorphic for every and and a finite subset Ξ in , where .
Proof. We have that a sequence is strictly increasing and satisfies the gap condition. This implies that . Without loss of generality, a set can also be ordered into a strictly increasing sequence.
By virtue of Theorem 9.1.6 [
20], the Müntz space
contains a complemented isomorphic copy of
; consequently,
and
are linearly topologically isomorphic as normed spaces.
Then, from Lemma 1 taking
, we deduce that:
and:
for each
, and hence,
is isomorphic with
. Considering the set
and then the set
, we get that
and
are linearly topologically isomorphic as normed spaces, as well. ☐
Theorem 1. Let increasing sequences and of positive numbers satisfy Conditions 1(1,2), and let for each n. If , where , then and are the isomorphic Banach spaces, where .
Proof. There exist the natural isometric linear embeddings of the Müntz spaces
and
into
. We choose a sequence of sets
satisfying the following restrictions (1)–(4):
- (1)
and for each and , where ;
- (2)
for each and ;
- (3)
is a monotone decreasing subsequence, which may be finite or infinite, having positive terms tending to zero. The terms are obtained from the sequence by elimination of zero terms. Denote by the corresponding enumeration mapping, such that for each is not zero;
- (4)
is a monotone increasing sequence with .
Take an arbitrary function
f in
. In view of Theorem 6.2.3 and Corollary 6.2.4 [
20], a function
f has a power series expansion:
where
for each
, where the power series decomposition of
f converges for each
, since
f is analytic on
.
Therefore, for each , we consider the power series
. Then, we infer that:
so that
is a monotone decreasing sequence by
l, and hence:
according to Dirichlet’s criterion (see, for example, [
25]) for each
, where
. Then, we deduce that:
since the mapping
is the orientation preserving the diffeomorphism of
onto itself, also
for each
by Lemma 7.3.1 [
20] and:
Thus, the series converges on , and the function is analytic on .
Inequality (3) implies that the linear isomorphism
of
with
exists, such that
,
. Then, we take the sequence of operators
. The space
is complete, and the operator sequence
converges relative to the operator norm to an operator
, so that
, since:
and
, where
I denotes the unit operator. Therefore, the operator
S is invertible. On the other hand, from Conditions (1)–(4), it follows that
. ☐
3. Approximation in Müntz Spaces
Now, we recall necessary definitions and notations of the Fourier approximation theory and then present useful lemmas.
Notation 2. Suppose that is a lower triangular infinite matrix with real matrix elements satisfying the restrictions: for each , where are nonnegative integers. To each one-periodic function in the space or in , is posed a trigonometric polynomial:where and are the Fourier coefficients of a function .
For measurable one-periodic functions
h and
g, their convolution is defined whenever it exists by the formula:
Putting the kernel of the operator
as:
we get:
The norms of these operators are:
which are constants of a summation method, where
denotes a norm on a Banach space
E, where either
or
with
, while
is a marked real number.
Henceforward, the Fourier summation methods prescribed by sequences of operators
that converge on
E:
in the
E norm will be considered.
Henceforth,
denotes the set of all pairs
satisfying the conditions:
is a sequence of non-zero numbers for which
the limit is zero,
is a real number and also:
is the Fourier series of some function from
. By
is denoted the family of all positive sequences
tending to zero with
for each
k so that the series:
converges. The set of all downward convex functions
for each
, so that
is denoted by
, while
is its subset of functions satisfying Condition (11).
Then:
is the approximation precision of
f by the Fourier series
, where:
is the partial Fourier sum approximating a Lebesgue integrable one-periodic function
f on
.
Definition 2. Suppose that and is its Fourier series with coefficients and , while is an arbitrary sequence that is real or complex. If the function:belongs to the space of all Lebesgue integrable (summable) functions on , then is called the Weil derivative of f. Then, stands for the family of all functions with ; we also put .
Particularly, for , this space is Weil–Nagy’s class , and the notation can be used instead of in this case. Put particularly , where .
Then, let
, where
X is a subset in
denotes the family of all trigonometric polynomials
of a degree not greater than
.
Lemma 3. Suppose that for each , where , , . Then, for each , there exists , such that the operator from into has the norm .
Proof. The Banach spaces and are defined with the help of the Lebesgue measure on . Then, Equation 2 implies that as soon as . That is, when , since and . ☐
Corollary 1. Let and , so that ; let also , where , while I is the unit operator. Then, is isomorphic with .
Proof. There is the natural embedding of
into
when
and
, such that
for each
, where
notates the characteristic function of a set
A. Since
, then the operator
is invertible (see [
26]).
Lemma 4. Let , where . Then:where .
Proof. Since
, then
is a
-additive and finite measure on
, where
is the Lebesgue measure on
(see, for example, [
27], Theorems V.5.4.3 and V.5.4.5 [
26]). Therefore, the limit exists:
☐
From Holder’s inequality, it follows that:
hence:
Thus, from Equations (14) and (15), the statement of this lemma follows. ☐
Note: We remind about the following definition: the family of all Lebesgue measurable functions
satisfying the condition:
is called the weak
space and denoted by
, where
notates the Lebesgue measure on the real field
,
,
(see, for example, §9.5 [
27], §IX.4 [
28,
29]).
The following Proposition 1 is used below in Theorem 2 to prove that functions of Müntz spaces for Λ satisfying the Müntz condition and the gap condition belong to Weil–Nagy’s class, where .
Proposition 1. Suppose that an increasing sequence of natural numbers satisfies the Müntz condition, and . Then, for a function , where .
Proof. In view of Theorem 6.2.3 and Corollary 6.2.4 [
20], a function
f is analytic on
, and consequently,
h is analytic on
; hence, a derivative
is also analytic on
. Moreover, the series:
converges on
, where
denotes the open disk in
of radius
with center at
, where
is an expansion coefficient for each
. That is, the functions
f and
h have holomorphic univalent extensions on
, since
(see Theorem 20.5 in [
30]). Take the function
, where
. In virtue of Theorem VI.4.2 [
26] and Lyapunov’s inequality (Equation
in §II.6 [
31]), this function is continuous, so that
. Together with Equation (1), this implies that the function
belongs to
and has a holomorphic univalent extension on
.
Then, we put
for each
, where
. From Lemma 4, it follows that:
Thus, the function is holomorphic (may be multivalent because of the multiplier ) on and continuous on .
According to Cauchy’s Equation 21
in [
30]:
for each
, where
is an oriented rectifiable boundary
of a simply-connected open domain
G contained in
, such that
. Particularly, this is valid for each
z in
and
.
On the other hand, the function
is bounded on
, where
notates the closed disk of radius
with the center at
. Thus,
. Estimating the integral (18) and taking into account Equation (17), we infer that
for each
, since
. Together with the analyticity of
on
this implies that:
where
. Thus,
. ☐
Theorem 2. Let an increasing sequence of natural numbers satisfy the Müntz condition, also and , and let , where . Then, for each , there exists , so that .
Proof. Let and , then is analytic on , since f is analytic on and . We take its one-periodic extension on .
According to Proposition 1.7.2 [
32] (or see [
33]),
if and only if there exists a function
, which is one-periodic on
, and Lebesgue integrable on
, such that:
where
(see Notation 2 and Definition 2).
We take a sequence
given by Equation (6), so that:
and write for short
instead of
. Under these conditions, the limit exists:
in
norm for each
according to Chapters 2 and 3 in [
32] (see also [
15,
33]).
On the other hand, Equation I
[
32] provides:
where
is the Fourier series corresponding to a function
, when
.
Put
for all
. Then,
for each
due to Theorems II.13.7, V.1.5 and V.2.24 [
33] (or see [
15]). This is also seen from Chapters I and V in [
32] and Equations (19) and (21) above. In view of Dirichlet’s theorem (see §430 in [
25]), the function
is continuous on the segment
for each
.
According to Equation 2.5.3.
in [
34]:
for each
and
. On the other hand, the integration by parts gives:
for every
,
and
. From Equation V
, Theorems V.2.22 and V.2.24 in [
33] (see also [
18,
35]), we infer the asymptotic expansions:
in a small neighborhood
of zero, where
,
,
and
are real constants. Taking
, we get that
Evidently, for Lebesgue measurable functions
and
, there is the equality
for each
whenever this integral exists, where
denotes the characteristic function of a subset
A in
, such that
for each
, also
for each
y outside
A,
. Particularly, if
, where
is a constant, then
(see also [
25,
26]). This is applicable to Equation
putting
there and with the help of the equality:
for each
and one-periodic functions
f and
g and using also that
for the considered types here of norms for each
, where
and
for each
, since:
We mention that according to the weak Young inequality:
for each
and
, where
,
and
,
is a constant independent of
and
(see Theorem 9.5.1 in [
27], §IX.4 in [
28]).
By virtue of Equation (21), the weak Young inequality (22) and Proposition 1, there exists a function
s in
, so that:
where
Therefore,
and
according to Equations (20) and (22). Thus,
. ☐
Below, Lemma 5 and Proposition 2 are given. They are used in Theorem 3 for proving the existence of a Schauder basis. On the other hand, Theorem 2 is utilized to prove Lemma 5.
Lemma 5. If an increasing sequence Λ
of natural numbers satisfies the Müntz condition, also and , :then a positive constant exists, so that:for each natural number .
Proof. By virtue of Theorem 2, the inclusion
is valid for each
, where
is in
with
for each
,
. Then,
for each
, since:
☐
Therefore, (see also §7), where for a linear space Y over and a marked real number b.
Then, the estimate (23) follows from Theorem V.5.3 in [
32].
4. Existence of Schauder Basis
Proposition 2. Let X be a Banach space over , and let Y be its Banach subspace, so that they fulfill the conditions –
below:- (1)
there is a sequence in X, such that are linearly independent vectors and for each n and;
- (2)
there exists a Schauder basis in X, such that:for each , where are real coefficients; - (3)
for every and , there exist , so that:where is a strictly monotone decreasing positive function with and, - (4)
where for each natural numbers k and l, where a sequence of normalized vectors in Y is such that its real linear span is everywhere dense in Y and and for each ; - (5)
vectors ,..., are linearly independent in Y for each .
Then, Y has a Schauder basis.
Proof. The real linear span
is complemented in
Y for each
due to Theorem
in [
4]. Put
and
, where
denotes the closure of a subset
A in
X, where
denotes the real linear span of
A. Since
Y is a Banach space and
for each
k, then
and
for each natural number
n and
m. Then, we infer that:
where
.
Take arbitrary vectors
and
, where
. Therefore, there are real coefficients
and
, such that:
and:
Hence, due to Condition (2):
and:
On the other hand:
consequently:
where
and
, when
.
When
and
, we infer using the triangle inequality that
for the best approximation
h of
in
, since
for each
j. Therefore, the inequality
and
imply that there exists
, such that the inclination of
to
is not less than
for each
. Condition (4) implies that
is complemented in
Y. By virtue of Theorem 1.2.3 [
20], the Schauder basis exists in
Y. ☐
Theorem 3. If a set Λ satisfies the Müntz and gap conditions and , then the Müntz space has a Schauder basis.
Proof. In view of Lemma 2 and Theorem 1, it is sufficient to prove the existence of a Schauder basis in the Müntz space for . We mention that if the Müntz space over the real field has the Schauder basis, then over the complex field has it as well. Thus, it is sufficient to consider the real field .
Let
be kernels of the Fourier summation method in
as in Notation 2, such that:
- (1)
.
For example, Cesaro’s summation method of order one can be taken, to which Fejér kernels
correspond, so that the limit:
converges in
(see Theorem 19.1 and Corollary 19.2 in [
36]). That is, there exists the Schauder basis
in
, such that:
and:
for every
and
, where
and
are real expansion coefficients.
By virtue of Theorem 6.2.3 and Corollary 6.2.4 [
20], each function
has an analytic extension on
, and hence:
- (2)
are the convergent series on the unit open disk
in
with the center at zero (see Proposition 1), where
and
,
,
,
for each
. On the other hand, the Müntz spaces
and
are isomorphic for each
(see Lemma 1 above). Therefore, we consider henceforward the Müntz space
on the segment
, where
. We mention that
and
are isomorphic (see Theorem 2). Then,
and
are isomorphic, as well. In view of Corollary 1, it is sufficient to prove the existence of a Schauder basis in
.
Take the finite dimensional subspace
in
, where
. Due to Lemma 2, the Banach space
exists and is isomorphic with
. By virtue of Equation I
[
32]
, where
, when
.
Consider the trigonometric polynomials for , where . Put as the completion of the linear span , where , , .
It is known (see Proposition 1.7.1 [
32]) that
if and only if there exists
, so that
, where the function
is prescribed by Equation (10); the constant
is as above. In view of Lemma 2, it is sufficient to consider the case
.
There exists a countable subset in , such that with for each and so that is dense in , since is separable. Using Properties (1) and (2) in this proof, Proposition 1 and Lemma 5, we deduce that a countable set K and a sufficiently large natural number exist, so that the Banach space is isomorphic with and , where and . Therefore, by the construction above, the Banach space is the completion of the real linear span of a countable family of trigonometric polynomials .
Without loss of generality, this family can be refined by induction, such that
is linearly independent of
over
for each
. With the help of transpositions in the sequence
, the normalization and the Gaussian exclusion algorithm, we construct a sequence
of trigonometric polynomials that are finite real linear combinations of the initial trigonometric polynomials
and satisfying the conditions:
- (3)
- (4)
the infinite matrix having the
l-th row of the form
for each
is upper trapezoidal (step), where:
with
and
, where
,
or
when
;
for each
and
.
Then, as X and Y in Proposition 2, we take and . In view of Proposition 2 and Lemma 2, the Schauder basis exists in and, consequently, in , as well. ☐