: the Faber-Schauder basis, the Riesz representation theorem, and the Borel -algebra
1 Introduction
In this note I work out some results about the space of continuous functions . In some cases these are instances of general results about continuous functions on compact Hausdorff spaces or compact metrizable spaces.
2
Let be a compact topological space (not necessarily Hausdorff) and let be the collection of continuous functions . With the norm
this is a real Banach algebra, with unity defined by for all .11 1 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 124, Lemma 3.97. Generally, if is a compact metrizable space and is a separable metrizable space then is separable.22 2 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 125, Lemma 3.99. Thus, when is a compact metrizable space, is a separable unital Banach algebra.
3 Schauder bases
If is a real normed space, a Schauder basis for is a sequence in such that for each there is a unique sequence of real numbers such that . A sequence in is called a basic sequence if it is a Schauder basis for the closure of its linear span.33 3 Because Schauder bases are not as familiar objects as orthonormal bases (Hilbert space bases) or vector space bases (Hamel bases), it is worth explicitly checking their properties to make ourselves familiar with how they work.
If is a Schauder basis, then for all . Suppose that for some real numbers we have . Then for , which means that the set is linearly independent. Therefore a Schauder basis is linearly independent. It is immediate that the linear span of a Schauder basis is a dense linear subspace of . The following shows that a normed space with a Schauder basis is separable.44 4 In particular, a Banach space with a Schauder basis is separable. There is a celebrated counterexample found by Per Enflo of a separable Banach space for which there does not exist a Schauder basis.
Lemma 1.
If is a Schauder basis for a normed space then
is dense in .
Proof.
Let and let . There is some for which
For each , let satisfy . Then
which proves the claim. ∎
For each we define by . For , and , so and therefore , and for , and therefore . This shows that is linear. It is apparent that .
We define by
and it is immediate that is linear and that for each , as . It is also immediate that is a finite-rank operator: is a finite-dimensional linear subspace of . For ,
showing that
In particular, , namely, is a projection operator. We prove that when is a Banach space, is continuous.55 5 N. L. Carothers, A Short Course in Banach Space Theory, p. 26, Theorem 3.1. We indicate explicitly in the proof when we use that is a Banach space rather than merely a normed space.
Theorem 2.
If is a Banach space and is a Schauder basis for , then each is continuous, and . Furthermore, each is continuous.
Proof.
For , , so , which implies that . It thus makes sense to define
It is immediate that and that . If , then for all , which implies that . Therefore is a norm on .
For and ,
showing that is a bounded linear operator with operator norm . Suppose that is a Cauchy sequence in the norm . Then we have for each that is a Cauchy sequence in the norm , and because is a Banach space, there is some such that in the norm as . For there is some such that when , , and thus for and for any ,
(1) |
Now, for any and any ,
so using the above, for ,
Because as , there is some such that when , and in this case for ,
which shows that is a Cauchy sequence in the norm , and because is a Banach space there is some such that in the norm .
For any , the restriction of the linear map to the finite-dimensional linear subspace is continuous in the norm , thus
Using this, we find by induction that for each ,
and because in the norm as and is a Schauder basis, this implies that for all , and thus . Therefore
For and , by (1) we have and therefore , which shows that in the norm as . Thus, is a Banach space.
The identity map is a linear isomorphism, and for ,
showing that is continuous . Because and are Banach spaces and is a continuous linear isomorphism, by the open mapping theorem66 6 Walter Rudin, Functional Analysis, second ed., p. 49, Corollary 2.12c. there is some such that
Then
which means that is a bounded linear operator with operator norm . ∎
Each is a bounded linear operator on , and for each , , which means that in the strong operator topology. The fact that the functions are linear and continuous means that they belong to the dual space .
In Theorem 2, if , we call the Schauder basis monotone.
The following theorem gives sufficient and necessary conditions under which a sequence in a Banach space is a basic sequence.77 7 Joseph Diestel, Sequences and Series in Banach Spaces, p. 36, Chapter V, Theorem 1. Thus if a sequence satisfies this condition and its linear span is dense in space, then it is a Schauder basis.
Theorem 3.
Suppose that is a Banach space and that is a sequence of nonzero elements of . There is some such that for any sequence of real numbers and any ,
(2) |
if and only if is a basic sequence.
Proof.
It is apparent that . Let be the linear span of , and let be the linear span of . For ,
(3) |
Thus if then using the above with , for we get , showing that is linearly independent. Because is linearly independent, it makes sense to define a linear map by
For , if then by (2), , and if then . Thus for each , is a bounded linear operator with operator norm , and because is a Banach space and , there is a unique bounded linear operator whose restriction to is equal to , which satisfies .88 8 Gert K. Pedersen, Analysis Now, revised printing, p. 47, Proposition 2.1.11. For , , and for , there is a sequence that tends to , thus
(4) |
For and , there is some with . Then there are some for which , and for ,
showing that . It follows from this and (4) that for each there is a sequence of real numbers such that . If is another such sequence, let , and then we obtain from (3) that for each . Therefore, is the unique sequence of real numbers such that , which establishes that is a Schauder basis for . ∎
4 Haar system and Faber-Schauder system
For , for , and for , write
and we write .99 9 We are partly following the presentation in B. S. Kashin and A. A. Saakyan, Orthogonal Series, p. 61, Chapter III. Thus
If , let and . Thus
and
Lemma 4.
If then either or .
Proof.
Let and , with . Then
There are three cases: (i) , (ii) , (iii) . In the first case, , i.e. . In the second case, and , so , i.e. . In the third case, so , i.e. . ∎
We define , and for , , and , we define
For example,
and
and
We call the Haar system. It is a fact that is a monotone Schauder basis for the Banach space with the norm .1010 10 Joram Lindenstrauss and Lior Tzafriri, Classical Banach Spaces I and II, p. 3.
Now we define and for we define by
Each belongs to , and we call the Faber-Schauder system. For example,
and
and
and
Generally for , for , and for ,
We remark that
5 Riesz representation theorem
Let be a measurable space. A signed measure is a function such that (i) , (ii) assumes at most one of the values , and (iii) if is a sequence of disjoint elements of then . A finite signed measure is a signed measure whose image is contained in . We denote by the collection of all finite signed measures on . For and for , define
for , and we check that with addition and scalar multiplication thus defined is a real vector space. A positive measure is a signed measure whose imaged is contained in . For , we write
if is a positive measure; in any case . We check that is a partial order on with which is an ordered vector space. Finally, a probability measure is a positive measure satisfying , which in particular belongs to .
For , a partition of is a countable subset of whose members are pairwise disjoint. For and we define
It is immediate that . It is proved that is a finite positive measure on , called the total variation measure of .1111 11 Walter Rudin, Real and Complex Analysis, third ed., pp. 117–118, Theorem 6.2 and Theorem 6.4. For , define
Because is a finite positive measure and , and are finite positive measures, called the positive and negative variations of . Then , called the Jordan decomposition of .
For , define
One checks that is a norm on .
For a compact Hausdorff space and for , write when for all . A positive linear functional is a linear map such that when . In this case, because ,
i.e. , and because ,
i.e. , showing because that the operator norm of is , and in particular that .
For normed spaces and , an isometric isomorphism from to is a linear isomorphism satisfying for all . The simplest version of the Riesz representation theorem is for compact metrizable spaces, for which we do not need to speak about regular Borel measures or continuous functions vanishing at infinity.1212 12 Walter Rudin, Real and Complex Analysis, third ed., p. 130, Theorem 6.19.
Theorem 5 (Riesz representation theorem for compact metrizable spaces).
Let be a compact metrizable space and define by
is an isometric isomorphism, and is order preserving: if then is a positive linear functional, and thus if then .
6 The Borel -algebra of
Let , with the relative topology inherited from , with which is a compact metric space. For , we define by
which is continuous.
For a set and a collection of functions , the coarsest -algebra on such that each is measurable , where has the Borel -algebra, is called the -algebra generated by , and is denoted by . We show that the Borel -algebra of is equal to the -algebra generated by the family of projection maps .1313 13 K. R. Parthasarathy, Probability Measures on Metric Spaces, p. 212, Theorem 2.1.
Theorem 6.
Let be the -algebra generated by the family . Then .
Proof.
Because is a separable metric space it is second-countable, and so if is an open subset of then is equal to the union of countably many open balls. Each open ball is equal to the union of countably many closed balls: . Therefore each open subset of is equal to the union of countably many closed balls, and to prove that it suffices to prove that all closed balls belong to . To this end, let be an enumeration of , let , and let . Suppose that for all , and take . Then there is a subsequence of that tends to , and because is continuous, , and then because for each we have it follows that . This establishes
But
which belongs to . Thus is a countable intersection of elements of and so belongs to , which shows that .
On the other hand, for each the map is continuous and hence is measurable .1414 14 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 140, Corollary 4.26. Therefore . ∎
7 Relatively compact subsets of
If is a metric space, a subset of is called totally bounded if for each there are finitely many points such that for any point there is some for which . It is immediate that a compact metric space is totally bounded, and the Heine-Borel theorem states that a metric space is compact if and only if it is complete and totally bounded.1515 15 Charalambos D. Aliprantis and Kim C. Border, Infinite Dimensional Analysis: A Hitchhiker’s Guide, third ed., p. 86, Theorem 3.28. On the other hand, one checks that if is a metric space and is a totally bounded subset of , then the closure is a totally bounded subset of . Thus, if is a totally bounded subset of a complete metric space , then the closure is itself a complete metric space (because is a complete metric space), and because is complete and totally bounded, by the Heine-Borel theorem it is compact.
If is a topological space and is a subset of , we say that is equicontinuous at if for each there is a neighborhood of such that for all and for all , and we call equicontinuous if it is equicontinuous at each . We call pointwise bounded if for each , is a bounded subset of . The Arzelà-Ascoli theorem states that for a compact Hausdorff space and for , is equicontinuous and pointwise bounded if and only if is totally bounded.1616 16 Gerald B. Folland, Real Analysis: Modern Techniques and Their Applications, second ed., p. 137, Theorem 4.43. Then the closure is totally bounded and is itself a complete metric space, and hence is a compact metric space. That is, for a compact Hausdorff space , is equicontinuous and pointwise bounded if and only if is relatively compact in .
For and , define
For because is continuous and is compact, is uniformly continuous on . Thus for there is some such that when , , hence if then , i.e. as . We give sufficient and necessary conditions for a subset of to be relatively compact.
Lemma 7.
Let . is relatively compact if and only if
(5) |
and
(6) |
Proof.
For and for , let
which is an open neighborhood of . The Arzelà-Ascoli theorem tells us that is relatively compact if and only if is equicontinuous and pointwise bounded. If is relatively compact, then being pointwise bounded yields (5). Let . Because is equicontinuous, for each there is some such that for all and for all we have . Then because is compact, there are such that . Let be the Lebesgue number for this open cover: for each there is some for which . For , for , and for , there is some for which , so , which means that and , and thus . Therefore , and this is true for all which proves (6).
Suppose now that (5) and (6) are true. By (6), there is some such that , and therefore for , . With (5) this yields , whence is pointwise bounded. Let and let . By (6) there is some such that . Thus for and for , , which shows that is equicontinuous at . This is true for all , so is equicontinuous and by the Arzelà-Ascoli theorem we get that is relatively compact in . ∎
8 Continuously differentiable functions
Let be a function. We say that is differentiable at if there is some such that1717 17 cf. Nicolas Bourbaki, Elements of Mathematics. Functions of a Real Variable: Elementary Theory, p. 3, Chapter I, §1, no. 1, Definition 1.
If is differentiable at each , we say that is differentiable on and define by . We define to be the collection of those such that is differentiable on and is continuous. is contained in , and it turns out that is a Borel set in .1818 18 Alexander S. Kechris, Classical Descriptive Set Theory, p. 70, §11.B, Example 2. On the other hand, the collection differentiable functions , which is a subset of , is not a Borel set in .1919 19 S. M. Srivastava, A Course on Borel Sets, p. 139, Proposition 4.2.7.