Functional Analysis
0. Preface
Functional Analysis can mean many different things to different mathematicians. The core of the subject, however, is to study linear spaces with some topology which allows us to do analysis; ones like spaces of functions, spaces of operators acting on the space of functions, etc.
The book consists of two parts. The first part covers the basics of Banach spaces theory with the emphasis on its applications. The second part covers topological vector spaces, especially locally convex ones, generalization of banach spaces. In both parts, we give principal results e.g., the closed graph theorem, resulting in some repetition. One reason for doing this is that one often only needs a Banach-version of such results. Another reason is that the approach seems more pedagogically sound; the statement of the results in their full generality may (at first) look cryptical to students new to them.
Knowledge of measure theory will not be needed except for Chapter 3.4 5 (Lp spaces). On the other hand, solid knowledge in general topology is mandatory.
1. Preliminaries
In this chapter we gather some standard results, some of which are not necessarily in functional analysis proper yet will be used later, primarily to fix language and formulation. We prefer to omit proofs for them since they can be easily found somewhere else.
1. Theorem (Ascoli) Let be a family of continuous real-valued functions on a compact space . If
- (i) For each , .
- (ii) For each , every point has a neighborhood such that for every and every (i.e., is equicontinuous),
then
- is totally bounded in the metric .
1. Theorem (Heine-Borel) A totally bounded subset of a complete metric space is relatively compact.
As a corollary, the conclusion of the Ascoli's theorem thus says the closure of is compact since is complete.
A subset of a topological space is said to be meager if it is a union of countably many subsets of the space that have dense complement.
1. Theorem (Baire) Every complete metric space is non-meager.
Proof: (TODO: we should provide one.)
By modifying the proof so to use compact sets instead of balls we can show that every locally compact (Hausdorff?) space is also non-meager, though we will not be needing this.
1. Exercise Show by contradiction that the set of real numbers is uncountable, using the theorem.
In particular, a dense set may be meager.
1. Exercise Give an example of a dense but meager set. (Hint: you can find such an example in the book.)
1. Theorem (Tychonoff) Every product space of a nonempty collection of compact spaces is compact.
An linear operator from a linear space to a scalar field is called a linear functional.
1. Theorem (Hahn-Banach) Let be a real vector space and be a function on such that and for any and any . If is a closed subspace and is a linear functional on such that , then admits a linear extension defined in such that .
Proof: First suppose that for some . By hypothesis we have:
- for all ,
which is equivalent to:
- .
Let be some number in between the sup and the inf. Define for . It follows that is an desired extension. Indeed, on being clear, we also have:
- if
and
- if .
Let be the collection of pairs where is linear space with and is a linear function on that extends and is dominated by . It can be shown that is partially ordered and the union of every totally ordered sub-collection of is in (TODO: need more details). Hence, by Zorn's Lemma, we can find the maximal element and by the early part of the proof we can show that .
We remark that a different choice of in the proof results in a different extension. Thus, an extension given by the Hahn-Banach theorem in general is not unique.
1. Exercise State the analog of the theorem for complex vector spaces and prove that this version can be reduced to the real version. (Hint: )
(TODO: mention moment problem.)
1 Lemma Let be linear functionals on the same linear space. is a linear combination of if and only if .
Proof: The direct part is clear. We shall show the converse by induction. Suppose . We may suppose that there is such that . For any , since ,
- or .
The basic case is thus proven. Now, suppose the lemma holds for some . As before, we may suppose that there is some such that
- while .
For any , since for every , . Hence, the application of the inductive hypothesis to the linear functional gives:
for some scalars . (TODO: the proof can be clearer and simpler).
Part 1: Banach spaces
2. Normed space
Let be a linear space. A norm is a real-valued function on , with the notation , such that
- (i) (w:triangular inequality)
- (ii) for any scalar
- (iii) implies .
(ii) implies that . This and (ii) then implies for all ; that is, norms are always non-negative. With the metric a normed space is a metric space. If only (i) and only (ii) hold, the function is called a seminorm.
A linear space with a norm is called a normed space. We define the operator norm of a continuous linear operator between normed spaces and , denoted by , by
2 Theorem Let be a linear operator from a normed space to a normed space .
- (i) is continuous if and only if there is a constant such that for all
- (ii) any as in (i) , and if has nonzero element.
Proof: See w:bounded operator and see w:operator norm.
A complete normed space is called a Banach space. While there is seemingly no prototypical example of a Banach space, we still give one example of a Banach space: , the space of all continuous functions on a compact space , can be identified with a Banach space by introducing the norm:
It is a routine exercise to check that this is indeed a norm. The completeness holds since, from real analysis, we know that a uniform limit of a sequence of continuous functions is continuous. In concrete spaces like this one, one can directly show the completeness. More often than that, however, we will see that the completeness is a necessary condition for some results (especially, reflexivity), and thus the space has to be complete. The matter will be picked up in the later chapter.
2 Theorem (completion) Every normed space can be identified with a dense subspace of some Banach space.
The completeness is essential, for example, in the next result.
2 Theorem Let be a normed linear space, and be a Banach space. Let be a subspace. If is dense, then a continuous linear operator admits a unique extension to .
Proof: For each , let be Cauchy such that . By continuity as . Since is complete, exists in , and we denote it by . From this definition It is clear that the extension is linear. Moreover, if there is another extension , then is the limit of ; whence, the extension is unique. Finally, is continuous since
- .
2 Theorem (open mapping theorem) Let be Banach spaces. If is a continuous linear surjection, then it is an open mapping; i.e., it maps open sets to open sets.
Proof: Let . Since is surjective, . Then by Baire's Theorem
A linear operator from a normed space to is said to be closed if its graph, that is the set , is closed in .
2 Corollary If is a continuous linear operator between Banach spaces with closed range, then there exists a such that if then for some with .
Proof: Since is an open mapping, there exists a such that
where by we denote an open ball of radius about . By replacing with we can also archive that
Let be given. Then there is with . Since , there exists such that . Let . Then
- and
Finally, if , just take .
2 Exercise Use a quotient map to simplify the proof. (Hint: the answer will be given in the chapter 4)
2 Corollary If and are Banach spaces, then the norms and are equivalent; i.e., each norm is dominated by the other.
Proof: Let be the identity map. Then we have:
- .
This is to say, is continuous. Since Cauchy sequences apparently converge in the norm , the open mapping theorem says that the inverse of is also continuous, which means explicitly:
- .
By the same argument we can show that is dominated by
The technique using the corollary will be exploited in the proof of the next result.
2 Theorem (closed graph theorem) Let be Banach spaces, and a linear operator. The following are equivalent.
- (i) is continuos.
- (ii) If and is convergent, then .
- (iii) The graph of is closed.
Proof: That (i) implies (ii) is clear. To show (iii), suppose is convergent in . Then converges to some or , and is convergent. Thus, if (ii) holds, . Finally, to prove (iii) (i), we note that Corollary 2.something gives the inequality:
since by hypothesis the norm in the left-hand side is complete. Hence, if , then .
Note that when the domain of a linear operator is not a Banach space (e.g., just dense in a Banach space), the condition (ii) is not sufficient for the graph of the operator to be closed. A function is said to be densely defined if its domain is dense.
2 Exercise Let be a closed densely defined operator and be a linear operator with . If there are constants such that (i) and and (ii) for every , then is a closed densely defined operator.
2 Corollary A linear functional is continuous if and only if it has closed kernel.
Proof: is closed if is continuous. To show the converse, suppose and is convergent. Since the corollary is obvious when is identically zero, we may suppose that there is such that . Then the sequence has a limit in the kernel of , since the kernel is closed. It follows:
- .
When are normed spaces, by we denote the space of all continuous linear operators from to .
2 Theorem If is complete, then every Cauchy sequence in converges to a limit and .
Proof: Let be a Cauchy sequence in operator norm. For each , since
and is complete, there is a limit to which converges. Define . is linear since the limit operations are linear. It is also continuous since . Finally, and as .
2 Theorem (uniform boundedness principle) Let be a family of continuos functions where is a normed linear space. Suppose that is non-meager and that:
- for each
It then follows: there is some open and such that
- (a)
If we assume in addition that each member of is a linear operator and is a normed linear space, then
- (b)
Proof: Let be a sequence. By hypothesis, and each is closed since is open by continuity. It then follows that some has an interior point ; otherwise, fails to be non-meager. Hence, (a) holds. To show (b), making additional assumptions, we can find an open ball . It then follows: for any and any with ,
- .
A family of linear operators is said to be equicontinuous if given any neighborhood of we can find a neighborhood of such that:
- for every
The conclusion of the theorem, therefore, means that the family satisfying the hypothesis of the theorem is equicontinuous.
2 Corollary Let be normed spaces. Let be a bilinear or sesquilinear operator. If is separately continuous (i.e., the function is continuous when all but one variables are fixed) and is complete, then is continuous.
Proof: For each ,
where the right-hand side is finite by continuity. Hence, the application of the principle of uniform boundedness to the family shows the family is equicontinuous. That is, there is such that:
- for every and every .
The theorem now follows since is a metric space.
Since scalar multiplication is a continuous operation in normed spaces, the corollary says, in particular, that every linear operator on finite dimensional normed spaces is continuous. The next is one more example of the techniques discussed so far.
2. Theorem (Hahn-Banach) Let be normed space and be a linear subspace. If is a linear functional continuous on , then there exists a continuos linear functional on such that on and .
Proof: Apply the Hahn-Banach stated in Chapter 1 with as a sublinear functional dominating . Then:
- ;
that is, .
2. Corollary Let be a subspace of a normed linear space . Then is in the closure of if and only if = 0 for any that vanishes on .
Proof: By continuity . Thus, if , then . Conversely, suppose . Then there is a such that for every . Define a linear functional for and scalars . For any , since ,
- .
Since the inequality holds for as well, is continuous. Hence, in view of the Hahn-Banach theorem, while we still have on and .
2 Corollary If is a normed space and for all , then
Proof: Suppose , and define for scalars . Now, is continuous since its domain is finite-dimensional, and so by the Hahn-Banach theorem . Thus, , or .
Here is a classic application.
2 Theorem Let be Banach spaces, be a linear operator. If implies that for every , then is continuous.
Proof: Suppose and . For every , by hypothesis and the continuity of ,
- .
Now, by the preceding corollary and the continuity follows from the closed graph theorem.
2 Theorem Let be a Banach space and . Then is bounded if and only if for every
Proof: By continuity,
- .
This proves the direct part. For the converse, define for . By hypothesis
- for every .
Thus, by the principle of uniform boundedness, there is such that:
- for every
Hence, in view of Theorem 2.something, for ,
- .
2. Corollary Let be Banach, and dense and linear. Then for every if and only if (1) and for every .
Proof: Since is Cauchy, it is bounded. This shows the direct part. To show the converse, let . If , then
By denseness, we can take so that .
2 Exercise Show the corollary may fail if is not bounded.
2 Theorem Let be a continuous linear operator into a Banach space. If where is the identity operator, then the inverse exists, is continuous and can be written by:
- for each in the range of .
Proof: For , we have:
- .
Since the series is geometric by hypothesis, the right-hand side is finite. Let . By the above, each time is fixed, is a Cauchy sequence and the assumed completeness implies that the sequence converges to the limit, which we denote by . Since for each , it follows from the principle of uniform boundedness that:
- .
Thus, by the continuity of norms,
- .
This shows that is a continuous linear operator since the linearity is easily checked. Finally,
- .
Hence, is the inverse to .
2 Corollary The space of invertible continuous linear operators is an open subspace of .
Proof: If and , then is invertible.
If is a scalar field and is a normed space, then is called a dual of and is denoted by . In view of Theorem 2.something, it is a Banach space.
3. Hilbert space
A Banach space is called a Hilbert space if for each ordered pair of elements in the space there is a unique complex (or real) number called an inner product of and , denoted by , subject to the following conditions:
- (i) The functional is linear.
- (ii)
- (iii) implies that
We define and this can be shown to be a norm. Indeed, it is clear that and (iii) is the reason that implies that . Finally, the triangular inequality follows from the next lemma.
3 Lemma (Schwarz's inequality) where the equality holds if and only if we can write for some scalar .
If we assume the lemma for a moment, since for any complex number it then follows:
Proof of Lemma: First suppose . If , it then follows:
where the equation becomes if and only if . Since we may suppose that , the general case follows easily.
3 Theorem A normed linear space is a pre-Hilbert space if and only if (polarization identity)
Proof: The direct part is clear. To show the converse, we define
- .
It is then immediate that , and . Moreover, since the calculation:
,
we have: . If is a real scalar and is a sequence of rational numbers converging to , then by continuity and the above, we get:
3 Lemma Let be a pre-Hilbert. Then in norm if and only if for any and as .
Proof: The direct part holds since:
- as .
Conversely, we have:
- as
3 Lemma Let be a convex closed subset of a Hilbert space. Then admits a unique element such that
- .
Proof: For each , there is some such that . That is, . Since is convex,
- and so .
It follows:
| as |
This is to say, is Cauchy. Since is a closed subset of a complete metric space, whence it is complete, there is a limit with . The uniqueness follows since if we have
where the right side is for the same reason as before.
3 Theorem (orthogonal decomposition) Let be a Hilbert space and be a subspace. For every we can write
where and , and and are uniquely determined by
Proof: Clearly is convex, and it is also closed since the image of a closed set under translation is again closed. Now, the application of the above lemma gives an element such that . Let . This turns out to be orthogonal to . Indeed, for any with , if , we have
since . That is, or . The uniqueness holds since if and and , then
- (since ).
The right-hand side is equal to 0 since . Thus, .
3 Corollary Let be a subspace of a Hilbert space . Then
- (i) if and only if is dense in .
- (ii) .
Proof: Left as an exercise.
3 Theorem (representation theorem) Every continuous linear functional on a Hilbert space has the form:
- with a unique and
Proof: Let . Since is continuous and any finite set is closed in or , is closed. If , then take . If not, by the above corollary, there is a nonzero orthogonal to . By replacing with we may suppose that . For any , since is in the kernel of and thus is orthogonal to , we have:
and so:
The uniqueness follows since for all means that . Finally, we have the identity:
where the last inequality is Schwarz's inequality.
3 Exercise Using Lemma 1.6 give an alternative proof of the preceding theorem.
In view of Theorem 3.something, for each , we can write: where , a closed subspace of , and . Denote each , which is uniquely determined by , by . The function then turns out to be a linear operator. Indeed, for given , we write:
- and
where and for . By the uniqueness of decomposition
- .
The similar reasoning shows that commutes with scalars. Now, for (where and ), we have:
That is, is continuous with . In particular, when is a nonzero space, there is with and and consequently . Such is called an orthogonal projection.
The next theorem gives an alternative to the Hahn-Banach theorem.
3 Theorem Let be a linear (not necessarily closed) subspace of a Hilbert space. Every continuous linear functional on can be extended to a unique continuous linear functional on that has the same norm and vanishes on
Proof: Define . By the same argument used in the proof of Theorem 2.something (Hahn-Banach) and the fact that , we obtain . Since on , it remains to show the uniqueness. For this, let be another extension with the desired properties. Since the kernel of is closed and thus contain , on . Hence, for any ,
- .
The extension is thus unique.
3 Theorem Let be an increasing sequence of closed subspaces, and be the closure of . If is an orthogonal projection onto , then for every . Proof: Let . Then is closed. Indeed, if and , then
and so . Since , the proof is complete.
3 Lemma (Bessel's inequality) If is a sequence orthonormal in a Hilbert space , then
- for any .
Proof: If , then . Thus,
- .
Letting completes the proof. .
2 Lemma A normed space is complete (in the sense of metric spaces) if and only if implies exists.
Proof: If , then since
- as and ,
exists by completeness. Conversely, suppose is a Cauchy sequence. Thus, for each , there exists an index such that for any . Let . Then . Hence, by assumption we can get the limit , and since
- as ,
we conclude that has a subsequence converging to ; thus, it converges to .
3 Theorem (Parseval) Let be a orthonormal sequence in a Hilbert space . Then the following are equivalent:
- (i) is dense in .
- (ii) For each , .
- (iii) For each , .
- (iv) (the Parseval equality).
Proof: Let . If , then it has the form: for some scalars . Since we can also write: . Let . Bessel's inequality and that is complete ensure that exists. Since
for all , we have , proving (i) (ii). Now (ii) (iii) follows since
- as
To get (iii) (iv), take . To show (iv) (i), suppose that (i) is false. Then there exists a with . Then
- .
Thus, (iv) is false.
3 Theorem (adjoint) Let be Hilbert spaces. Let be a linear operator from a subspace of to . For each , is continuous for all in the domain of if and only if there exists some such that:
- for all in the domain of .
Here is uniquely determined by if and only if the domain of is dense.
Proof: The existence follows from the Hahn-Banach theorem (or the projection argument for Hilbert spaces) and the Riesz representation theorem. (FIXME: need more detail). Suppose the uniqueness of . It then follows: if and , then:
and the uniqueness shows that and . Conversely, if is dense, then
- for implies that .
When the domain of is dense, for each satisfying the continuity condition, we denote in the theorem by . can be shown to be linear (since is densely defined but not because of linearlity), and we call it the adjoint of .
3 Theorem Let be Hilbert spaces and be a closed densely defined (i.e., having dense domain and closed graph) operator. Then:
- (i) is a closed densely defined.
- (ii) .
Proof: Let . If is orthogonal to the graph of , then
- for all in the domain of ;
that is, . Thus, :
- for .
This is to say: and . (ii) (i) implies the existence of and the equality holds since the following are equivalent:
- (a)
- (b) for .
- (c) for .
- (d) .
Note that the equivalence of (b) and (c) follows, as in the proof of (i), from that is in the domain of and so . We conclude that , and (by applying (ii)) that has closed graph.
3 Corollary Let be Hilbert spaces, and if and only if there is some such that:
- for every
Proof: If is defined, then take . As for the converse, the estimate means that the map is continuous; hence, .
3 Theorem Let be Hilbert spaces, and densely defined linear operators. Suppose . Then where the equality holds if
Proof: Let be given. Then
- for every . But, by definition, denotes an element such as .
Hence, in . Finally, if , then, by the fact we have just proved
- .
3 Theorem Let be a closed linear subspace of a Hilbert space . is an orthogonal projection onto if and only if and the range of is .
Proof: The direct part is clear except for . To see this, note that for any is a real number since . Hence,
- for every .
By Lemma 3.something . For the converse we only have to verify . But it holds since and for any .
As an application we shall prove:
3 Theorem Let be Hilbert spaces, be a closed densely defined linear operator. Then is surjective if and only if there is a such that
- for every .
Proof: Suppose is surjective. It suffices to show that
is bounded. If , then we can write and so:
- for every .
That is bounded now follows from Theorem 2.something. To show the converse, let be given. Since is injective, we can define a linear functional by for . For every ,
- .
Thus, is continuous on the range of . It follows from the Hahn-Banach theorem that we may assume that is defined and continuous on . Thus, by Theorem 3.something, we can write in . Then:
- for every ,
and so .
3 Corollary Let be as given in the preceding theorem. Then is closed if and only if is closed.
Proof: Suppose is closed. Since restricted to has the same range as , we only have to show this operator has the closed range. Thus, suppose is convergent and . Applying the preceding theorem with , we get:
- as .
Thus, there is a limit of and since is a closed operator, . Hence, is closed. The converse holds since .
3 Lemma Let be linear operators from a Hilbert space over to itself. If for , then .
Proof: Let . We have and . Summing the two we get: for . Taking gives for all or .
3 Lemma Let be a continuous linear operator on a Hilbert space . Then if and only if for all .
Continuous linear operators with the above equivalent conditions are said to be normal. For example, an orthogonal projection is normal. See w:normal operator for additional examples and the proof of the above lemma.
3 Theorem Let be a normal operator. If and are distinct eigenvalues of , then the respective eigenspaces of and are orthogonal to each other.
Proof: Let be the identity operator, and be arbitrary eigenvectors for , respectively. Since the adjoint of is , we have:
- .
That is, , and we thus have:
If is nonzero, we must have .
A sequence in a Hilbert space is said to converge weakly to if
3 Theorem Let be an orthogonal sequence in a Hilbert space . The series converges if and only if the series converges weakly. Proof: (TODO: write one using the uniform boundedness principle and completeness.)
3 Exercise Let be a Hilbert space with orthogonal basis , and be a sequence with . Prove that there is a subsequence of that converges weakly to some and that . (Hint: Since is bounded, by Cantor's diagonal argument, we can find a sequence such that is convergent for every .)
We will obtain a more general version of this in the next chapter.
4 Geometry of Banach spaces
The major topic of the chapter is the notion of reflexibility of Banach spaces, which, as it turns out, is a geometric property.
If is a Banach space, we can introduce a topology called the "weak topology" in addition to the original norm one, as follows.
From the definitions follows immediately:
4. Corollary
- (i) converges weakly (i.e., in the terms of a weak topology) to in if and only if converges to for every .
- (ii) converges weak-* to in if and only if converges to for every .
In particular, when is a Hilbert space, in view of Theorem 3.something the above discussion remains valid if we replace each occurrence of by for some .
We remark this easy-to-remember characterization does not sufficiently tell the whole story.
4. Corollary The continuous dual in terms of a weak topology is the same as the continuous dual in terms of a norm topology.
Proof: Let be the weak topology and be the norm topology. By definition and so: . Conversely, if , then is norm-continuous; thus, .
4. Theorem Let be a Banach space. A linear functional is continuous on if and only if it is continuous on the closed unit ball of
4 Theorem Every finite dimensional Banach space is reflexive (thus, complete).
Proof: (TODO)
A linear operator is said to be a compact operator if the image of the open unit ball under is relatively compact. We recall that if a linear operator between normed spaces maps bounded sets to bounded sets, then it is continuous. Thus, every compact operator is continuous.
4. Theorem Let be a reflexive Banach space and be a Banach space. Then a linear operator is a compact operator if and only if sends weakly convergent sequence to norm convergent ones.
Proof: [1] Let converges weakly to , and suppose is not convergent. That is, there is an such that for infinitely many . Denote this subsequence by . By hypothesis we can then show (TODO: do this indeed) that it contains a subsequence such that converges in norm, which is a contradiction. To show the converse, let be a bounded set. Then since is reflexive every countable subset of contains a sequence that is Cauchy in the weak topology and so by the hypothesis is a Cauchy sequence in norm. Thus, is contained in a compact subset of .
4 Lemma Let . A normed space is finite-dimensional if and only if its closed ball of radius is compact.
Proof: If is not finite dimensional, using w:Riesz's_lemma, we can construct a sequence such that:
- for any sequence of scalars .
Thus, in particular, for all . (TODO: fill gaps)
4 Corollary
- (i) Every finite-rank linear operator (i.e., a linear operator with finite-dimensional range) is a compact operator.
- (ii) Every linear operator with the finite-dimensional domain is continuous.
Proof: (i) is clear, and (ii) follows from (i) since the range of a linear operator has dimension less than that of the domain.
4 Theorem The set of all compact operators into a Banach space forms a closed subspace of the set of all continuous linear operators in operator norm.
Proof: Let be a linear operator and be the open unit ball in the domain of . If is compact, then is bounded (try scalar multiplication); thus, is continuous. Since the sum of two compacts sets is again compact, the sum of two compact operators is again compact. For the similar reason, is compact for any scalar . We conclude that the set of all compact operators, which we denote by , forms a subspace of continuous linear operators. To show the closedness, suppose is in the closure of . Let be given. Then there is some compact operator such that . Also, since is a compact operator, we can cover by a finite number of open balls of radius centered at , respectively. It then follows: for , we can find some so that and so . This is to say, is totally bounded and since the completeness its closure is compact.
4. Corollary If is a sequence of compact operators which converges in operator norm, then its limit is a compact operator.
4 Theorem (transpose) Let be Banach spaces, and be a continuous linear operator. Define by the identity . Then is continuous both in operator norm and the weak-* topology, and .
Proof: For any
Thus, and is continuous in operator norm. To show the opposite inequality, let be given. Then there is with . Using the Hahn-Banach theorem we can also find and . Hence,
- .
We conclude . To show weak-* continuity let be a neighborhood of in ; that is, for some . If we let , then
since . This is to say, is weak-* continuous.
4 Exercise State and prove the analog of Theorem 3.something using the transpose instead of a Hilbert adjoint operator.
5 Lp spaces
5 Theorem (Minkowski's inequality) If and , then:
- .
Proof (See Chapter 2 of An Introduction to Analysis for more details): The theorem is a consequence of Hölder's inequality:
- (where we let by )
which is then a simple consequence of the following inequality:
- (where a, b > 0)
To simplify notation we denote by and , respectively. First we have:
Then Hölder's inequality followed by division gives:
where . The desired inequality follows after noting and .
By letting be a counting measure we also obtain the analog for the series:
5 Corollary
5 Theorem Let , , and a linear subspace. If is closed in both and , then the topology for inherited from and coincide.
Proof: The identity map is continuous. Since is a bijection between Banach spaces, the open mapping theorem says is a homeomorphism.
(Remark: the theorem is just an exercise 2.something.)
5 Theorem spaces are uniformly convex (thus, complete and reflexive).
Proof: (TODO)
5 Theorem (lifting property of ) Let be Banach spaces, and be a linear surjection. If is a a continuous linear operator, there exists a continuous linear operator such that .
Proof: Let be a canonical basis for . Using Corollary 2.something to the open mapping theorem we obtain a sequence and constant so that:
- , and .
Then . Define by . Then , and is continuous since .
6 Banach algebras
Part 2: Topological vector spaces
Topological vector space
A vector space endowed by a topology that makes translations (i.e., ) and dilations (i.e., ) continuous is called a topological vector space or TVS for short.
A subset of a TVS is said to be:
- bounded if for every neighborhood of there exist such that for every
- balanced if for every scalar with
- convex if for any and any with .
1 Corollary for any if and only if is convex.
Proof: Supposing we obtain for all . Conversely, if is convex,
- , or for any .
Since holds in general, the proof is complete.
Define for scalars , vectors . If is a balanced set, for any , by continuity,
- .
Hence, the closure of a balanced set is again balanced. In the similar manner, if is convex, for
- ,
meaning the closure of a convex set is again convex. Here the first equality holds since is injective if . Moreover, the interior of , denoted by , is also convex. Indeed, for with
- ,
and since the left-hand side is open it is contained in . Finally, a subspace of a TVS is a subset that is simultaneously a linear subspace and a topological subspace. Let be a subspace of a TVS. Then is a topological subspace, and it is stable under scalar multiplication, as shown by the argument similar to the above. Let . If is a subspace of a TVS, by continuity and linearity,
- .
Hence, is a linear subspace. We conclude that the closure of a subspace is a subspace.
Let be a neighborhood of . By continuity there exists a and a neighborhood of such that:
It follows that the set is a union of open sets, contained in and is balanced. In other words, every TVS admits a local base consisting of balanced sets.
1 Theorem Let be a TVS, and . The following are equivalent.
- (i) is bounded.
- (ii) Every countable subset of is bounded.
- (iii) for every balanced neighborhood of there exists a such that .
Proof: That (i) implies (ii) is clear. If (iii) is false, there exists a balanced neighborhood such that for every . That is, there is a unbounded sequence in . Finally, to show that (iii) implies (i), let be a neighborhood of 0, and be a balanced open set with . Choose so that , using the hypothesis. Then for any , we have:
1 Corollary Every Cauchy sequence and every compact set in a TVS are bounded.
Proof: If the set is not bounded, it contains a sequence that is not Cauchy and does not have a convergent subsequence.
1 Lemma Let be a linear operator between TVSs. If is bounded for some neighborhood of , then is continuous.
1 Theorem Let be a linear functional on a TVS .
- (i) has either closed or dense kernel.
- (ii) is continuous if and only if is closed.
Proof: To show (i), suppose the kernel of is not closed. That means: there is a which is in the closure of but . For any , is in the kernel of . This is to say, every element of is a linear combination of and some other element in . Thus, is dense. (ii) If is continuous, is closed. Conversely, suppose is closed. Since is continuous when is identically zero, suppose there is a point with . Then there is a balanced neighborhood of such that . It then follows that . Indeed, suppose . Then
- if , which is a contradiction.
The continuity of now follows from the lemma.
1 Lemma Let be a sequence of subsets of a a linear space containing such that for every . If and , then .
Proof: We shall prove the lemma by induction over . The basic case holds since for every . Thus, assume that the lemma has been proven until . First, suppose are not all distinct. By permutation, we may then assume that . It then follows:
- and .
The inductive hypothesis now gives: . Next, suppose are all distinct. Again by permutation, we may assume that . Since no carry-over occurs then and , and so:
- .
Hence, by inductive hypothesis, .
1 Theorem Let be a TVS.
- (i) If is Hausdorff and has a countable local base, is metrizable with the metric such that
- and for every
- (ii) For every neighborhood of , there is a continuous function such that
- , on and for any .
Proof: To show (ii), let be a sequence of neighborhoods of satisfying the condition in the lemma and . Define on and for every . To show the triangular inequality, we may assume that and are both , and thus suppose and . Then
Thus, . Taking inf over all such we obtain:
and do the same for the rest we conclude . This proves (ii) since is continuous at and it is then continuous everywhere by the triangular inequality. Now, to show (i), choose a sequence of balanced sets that is a local base, satisfies the condition in the lemma and is such that . As above, define for each . For the same reason as before, the triangular inequality holds. Clearly, . If , then there are such that and . Thus, by the lemma. In particular, if for "every" , then since is Hausdorff. Since are balanced, if ,
- for every with .
That means , and in particular . Defining will complete the proof of (i). In fact, the properties of we have collected shows the function is a metric with the desired properties. The lemma then shows that given any , for some . That is, the sets over forms a local base for the original topology.
The second property of in (i) implies that open ball about the origin in terms of this is balanced, and when has a countable local base consisting of convex sets it can be strengthened to:, which implies open balls about the origin are convex. Indeed, if , and if and with , then
since the sum of convex sets is again convex. This is to say,
and by iteration and continuity it can be shown that for every .
Corollary For every neighborhood of some point , there is a neighborhood of with
Proof: Since we may assume that , take .
Corollary If every finite set of a TVS is closed, is Hausdorff.
Proof: Let be given. By the preceding corollary we find an open set containing .
Locally convex spaces
A TVS with a local base consisting of convex sets is said to be locally convex. Since in this book we will never study non-Hausdorff locally convex spaces, we shall assume tacitly that every finite subset of every locally convex is closed, hence Hausdorff in view of Theorem something.
Lemma Let be locally convex. The convex hull of a bounded set is bounded.
Distributions
The support of , denoted by , is the complement of the union of all the open sets such that for every .
Theorem Let . Then for every
Theorem A distribution has compact support if and only if it can be extended to an continuous linear functional on .
Theorem If , then
Proof: When , the theorem is a special case of Theorem something. Using Theorem something we can find a sequence such that in -norm. It follows:
- .
Hence, there is a limit such that in -norm and .
Template:Subject Template:Alphabetical Template:Shelf Template:DDC Template:LOC