The weak topology of locally convex spaces and the weak-* topology of their duals

Jordan Bell
April 3, 2014

1 Introduction

These notes give a summary of results that everyone who does work in functional analysis should know about the weak topology on locally convex topological vector spaces and the weak-* topology on their dual spaces. The most striking of the results we prove is Theorem 9, which shows that a subset of a locally convex space is bounded if and only if it is weakly bounded. It is straightforward to prove that if a set is bounded then it is weakly bounded, but to prove that if a set is weakly bounded then it is bounded we use the Hahn-Banach separation theorem, the Banach-Alaoglu theorem, and the uniform boundedness principle.

If X is a topological vector space then we will see that the weak topology on it is coarser than the original topology: any set that is open in the original topology is open in the weak topology. From this it follows that it is easier for a sequence to converge in the weak topology than in the original topology: for a sequence to converge to a point means that it is eventually contained in every neighborhood of the point, and a point has fewer neighborhoods in the weak topology than it does in the original topology. The weak topology encodes information we may care about, and we may be able to establish that certain sets are compact in the weak topology that are not compact in the original topology.

In these notes I first define the weak topology on a topological vector space X, and show that if X is locally convex then X with the weak topology is also a locally convex space. Indeed a normed space is locally convex, but there are function spaces that we care about that are not normed spaces. For example, the set of holomorphic functions on the open unit disc is a Fréchet space11 1 A Fréchet space is a complete metrizable locally convex space. It is a fact that a locally convex space is metrizable if and only its topology is induced by countably many of its seminorms. See John B. Conway, A Course in Functional Analysis, second ed., p. 105, chapter IV, Proposition 2.1. that is not normable. If U is an open subset of n, then Ck(U) is a Fréchet space that is not normable, and the set of Schwartz functions on U is also a Fréchet space that is not normable. Moreover, none of the theorems stated for topological vector spaces and locally convex spaces is much easier to prove in the case of normed spaces, and thus it is not a great waste of time to digest their statements in a larger category of spaces.

Except for the results about normed spaces, the hypotheses of all the theorems that we present are satisfied for a separable metrizable locally convex space. The hypotheses of every result in this note are satisfied for separable reflexive Banach spaces, for example, Lp(n) for 1<p<.

If every instance of in this text is replaced with , the resulting text does not require any further changes to be correct.

2 Topological vector spaces

If X is a topological space and xX and is a set of open sets, we say that is a local basis at x if each element of includes x and if for every open set U that includes x there is some V such that VU. If for each xX the set x is a local basis at x, then xXx is a basis for the topology of X. If X is a vector space, is a set of subsets of X, and xX, we define

x+={x+N:N},x+N={x+y:yN}.

If X is a vector space over with a topology 𝒪(X) such that (X,𝒪(X)) is Hausdorff and such that addition X×XX and scalar multiplication ×XX are continuous, we say that X is a topological vector space. It is straightforward to prove that if is a local basis at 0 then x+ is a local basis at x, and so xX{x+} is a basis for the topology 𝒪(X). To specify a topology on a vector space it suffices to specify a local basis at 0: This gives a basis by taking the union of the translates of the local basis over all xX, and then this basis generates a topology. However, X might not be a topological vector space with the topology thus generated. (That is, if we define a topology on a vector space by declaring certain sets including the origin to be open, the vector space need not be a topological vector space with this topology.)

A topological vector space (X,𝒪(X)) is said to be locally convex if there is a local basis at 0 each element of which is convex. If να is a family of seminorms on X, we define the seminorm topology induced by this family to be the coarsest topology on X such that for all x0X and αI, the map xνα(x-x0) is continuous. We say that a set of seminorms να is a separating family if x0 implies that there is some α such that να(x)>0. One can prove that a topological vector space is locally convex if and only if its topology is induced by a separating family of seminorms.22 2 Walter Rudin, Functional Analysis, second ed., p. 27, Theorem 1.36 and Theorem 1.37.

Let (X,𝒪(X)) be a topological vector space over . The dual space X* is the set of all continuous linear maps (X,𝒪(X)). The weak topology on X, which we denote by 𝒪w(X), is the initial topology for X*. That is, 𝒪w(X) is the coarsest topology on X so that each element of X* is continuous (X,𝒪w(X)). Equivalently, the weak topology on X is the seminorm topology induced by the seminorms |α|, αX*. The topologies 𝒪(X) and 𝒪w(X) are comparable, and 𝒪(X) is at least as fine as 𝒪w(X). That is, 𝒪w(X)𝒪(X). A vague rule is that the smaller X* is compared to the set of all linear maps X, the smaller 𝒪w(X) will be compared to 𝒪(X). If X* separates X then (X,𝒪w(X)) is a locally convex topological vector space. It is locally convex because 𝒪w(X) is induced by the separating family of seminorms |α|, αX*.

  • We say that xkx weakly if xkx in (X,𝒪w(X)): for every neighborhood N of x in (X,𝒪w(X)), the sequence xk is eventually in N.

  • We say that A is weakly bounded if A is a bounded subset of (X,𝒪w(X)): for every neighborhood N of 0 in (X,𝒪w(X)) there is some c0 such that A{cx:xN}=cN.

3 Locally convex spaces

The Hahn-Banach theorem separation theorem states the following.33 3 Walter Rudin, Functional Analysis, second ed., p. 59, Theorem 3.4.

Theorem 1 (Hahn-Banach separation theorem).

If

  • X is a locally convex topological vector space over

  • A,B are disjoint, nonempty, closed, convex subsets of X

  • A is compact

then there is some λX* and γ1,γ2R such that, for all aA and bB,

Reλ(a)<γ1<γ2<Reλ(b).

It follows that if X is locally convex then X* separates X. We can also use the Hahn-Banach separation theorem to prove that in a locally convex space, the weak closure of a convex set is equal to its original closure.44 4 Walter Rudin, Functional Analysis, second ed., p. 66, Theorem 3.12.

Theorem 2.

If E is a convex subset of a locally convex space (X,O(X)), then the closure E¯ of E in (X,O(X)) is equal to the closure E¯w of E in (X,Ow(X)).

Proof.

For AX, denote Ac=XA. We have, as 𝒪w(X) is coarser than 𝒪(X),

E¯ = (U𝒪(X),EUcU)c
(U𝒪w(X),EUcU)c
= E¯w.

In the other direction, let x0E¯. If we can show that x0E¯w, this will show that E¯wE¯ and hence that E¯=E¯w. Let A={x0} and B=E¯, which satisfy the conditions of the Hahn-Banach theorem. Thus there is some λX* and γ1,γ2 such that for all xE¯,

Reλ(x0)<γ1<γ2<Reλ(x).

Let U={xX:Reλ(x)<γ1}. The set V={z:Rez<γ1} is an open subset of and U=λ-1(V), so U𝒪w(X). For every xE¯ we have Reλ(x)>γ2>γ1, and for every xU we have Reλ(x)<γ1, so E¯U= and in particular EU=. The three facts x0U,U𝒪w(X), and EU= imply that x0E¯w, completing the proof. ∎

If a sequence converges weakly, it need not converge in the original topology. Mazur’s theorem shows that if a sequence in a metrizable locally convex space converges weakly then there is a sequence in the convex hull of the original sequence that converges to the same limit as the weak limit of the original sequence.55 5 Walter Rudin, Functional Analysis, second ed., p. 67, Theorem 3.13.

Theorem 3 (Mazur’s theorem).

Let X be a metrizable locally convex space. If xnx weakly, then there is a sequence yiX such that each yi is a convex combination of finitely many xn and such that yix.

Proof.

The convex hull of a subset A of X is the set of all convex combinations of finitely many elements of A. The convex hull of a set is convex and contains the set. Let H be the convex hull of the sequence xn, and let K be the weak closure of H. Since xnx weakly and xnH, it follows that xK. As H is convex, Theorem 2 tells us that K=H¯, so xH¯. But X is metrizable, so x being in the closure of H implies that there is a sequence yiH such that yix. This sequence yi satisfies the claim. ∎

4 Weak-* topology

If X is a vector space over and is a set of linear maps X that separates X, and we give X the initial topology for , then one can prove that with this topology X is a locally convex space whose dual space is .66 6 Walter Rudin, Functional Analysis, second ed, p. 64, Theorem 3.10.

Let X be a topological vector space over , and for xX, define fx:X* by fx(λ)=λ(x). fx is linear. If λ1,λ2X* are distinct, then λ1-λ20 so there is some xX such that (λ1-λ2)(x)0, which tells us that fx(λ1)fx(λ2). Therefore the set {fx:xX} separates X*. Let 𝒪X(X*) be the initial topology for {fx:xX}, and by the previous paragraph we have that (X*,𝒪X(X*)) is a locally convex space whose dual space is {fx:xX}. The topology 𝒪X(X*) is called the weak-* topology on X*. We record what we’ve just said as a theorem to make it easier to look up.

Theorem 4.

If X is a topological vector space, then its dual X* with the weak-* topology OX(X*) is a locally convex space, and the dual space of (X*,OX(X*)) is the set of fx:X*C, where fx(λ)=λ(x), xX,λX*.

The Banach-Alaoglu theorem77 7 Walter Rudin, Functional Analysis, second ed., p. 68, Theorem 3.15. shows that certain subsets of X* are weak-* compact, i.e. they are compact subsets of (X,𝒪X(X*)). The set K in the statement of the theorem is called the polar of the set V.

Theorem 5 (Banach-Alaoglu theorem).

If X be a topological vector space and VX is a neighborhood of 0, then

K={λX*:if xV then |λ(x)|1}

is a compact subset of (X*,OX(X*)).

The following theorem shows that in a separable topological vector space, a weak-* compact set is weak-* metrizable.88 8 Walter Rudin, Functional Analysis, second ed., p. 70, Theorem 3.16. This is useful because if possible we would like to characterize a topology by its convergent sequences rather than by its open sets.

Theorem 6.

If X is a separable topological vector space and K is a weak-* compact subset of X*, then K with the subspace topology inherited from (X*,OX(X*)) is metrizable.

Proof.

Let {xn} be a countable dense set in X, and define fn:X* by fn(λ)=λ(xn). For each n, fn:(X*,𝒪X(X*)) is linear and continuous. For distinct λ1,λ2X*, the set U={xX:λ1(x)λ2(x)} is an open subset of X. As U is a nonempty open set and {xn} is a dense subset of X, there is some xnU, giving fn(λ1)fn(λ2). Therefore {fn:X*} separates X*. It is a fact that if Y is a compact topological space and there is a countable set of continuous functions Y that separates Y then Y is metrizable.99 9 Walter Rudin, Functional Analysis, second ed., p. 63, §3.8, (c). In Rudin it is stated for real valued functions. K is a compact topological space with the subspace topology inherited from 𝒪X(X*), and hence this topology on K is metrizable. ∎

We can combine Theorem 5 and Theorem 6 to get the following, which states that the polar of a neighborhood of 0 in a separable topological vector space is weak-* sequentially compact.1010 10 Walter Rudin, Functional Analysis, second ed., p. 70, Theorem 3.17.

Theorem 7.

Let X is a separable topological vector space. If V is a neighborhood of 0 and if the sequence λnX* satisfies

|λn(x)|1,n1,xV,

then there is a subsequence λa(n) and some λX* such that for all xX,

limnλa(n)(x)=λ(x).
Proof.

The Banach-Alaoglu Theorem implies that the polar

K={λX*:if xV then |λ(x)|1},

is weak-* compact. K with the subspace topology inherited from 𝒪X(X*) is compact, hence by Theorem 6 it is metrizable. Since the sequence λn is contained in K, it has a subsequence λa(n) that converges weakly to some λK. For each xX, the function fx:(X*,𝒪X(X*)) defined by fx(λ)=λ(x) is continuous, hence for all xX we have fx(λa(n))fx(λ), which is the claim. ∎

The following lemma gives a tractable characterization of weakly bounded sets.1111 11 Walter Rudin, Functional Analysis, second ed., p. 66.

Lemma 8.

If X is a topological vector space and E is a subset of X, then E is weakly bounded if and only if for all λX* there is some γ(λ) such that if xE then |λ(x)|γ(λ).

The following theorem1212 12 Walter Rudin, Functional Analysis, second ed., p. 70, Theorem 3.18. shows that in a locally convex space, the boundedness of a set is equivalent to its weak boundedness. Its proof involves much of what we have talked about so far, and actually requires introducing some new terms with which one should probably have at least an acquaintance. The statement of the theorem does not involve the weak-* topology on X*, but the proof uses the Banach-Alaoglu theorem.

Theorem 9 (Weak boundedness is equivalent to boundedness).

If X is locally convex and EX, then E is bounded in (X,O(X)) if and only if E is bounded in (X,Ow(X)).

Proof.

Suppose that E is bounded in (X,𝒪(X)) (for every neighborhood of 0, the set E is contained in some dilation of that neighborhood). Let N be a neighborhood of 0 in 𝒪w(X). There is some U𝒪w(X) with 0UN. Now, U𝒪(X), so N is a neighborhood of x in 𝒪(X). As E is bounded, there is some α such that EαN={αx:xN}, which is what it means for E to be bounded in (X,𝒪w(X)).

Suppose that E is bounded in (X,𝒪w(X)). Let N be a neighborhood of 0 in (X,𝒪(X)). We want to show that there is some α such that E{αx:xN}. If is a local basis at a point in a topological vector space, then every element of contains the closure of some element of .1313 13 Walter Rudin, Functional Analysis, second ed., p. 11, Theorem 1.11. A subset A of a vector space is said to be balanced if for every α with |α|1 we have αAA. It is a fact that in a topological vector space, every convex neighborhood of 0 contains a convex balanced neighborhood of 0.1414 14 Walter Rudin, Functional Analysis, second ed., p. 12, Theorem 1.14. The purpose of having said all of this is the following: since N is a neighborhood of 0 in 𝒪(X), there is a balanced convex V that is a neighborhood of 0 in (X,𝒪(X)) such that V¯N, where V¯ is the closure of V in (X,𝒪(X)). Moreover, it is a fact that the closure of a balanced set is itself balanced, and the closure of a convex set is itself convex,1515 15 Walter Rudin, Functional Analysis, second ed., p. 11, Theorem 1.13. so V¯ is a balanced convex set that is contained in N. Thus to prove EαN it suffices to prove that EαV¯, and this is what we will do.

Let

K={λX*:if xV then |λ(x)|1},

the polar of V. If xV and λK then |λ(x)|1, so

V{xX:if λK then |λ(x)|1}=λK{xX:|λ(x)|1}.

The right-hand side is an intersection of closed sets in (X,𝒪(X)), so it is closed. Therefore V¯ is contained in the right-hand side. Furthermore, it is a consequence of the Hahn-Banach separation theorem1616 16 Walter Rudin, Functional Analysis, second ed., p. 61, Theorem 3.7. that if B is a convex balanced closed set in a locally convex space and x0B, then there is some λX* such that λ(x0)>1 and |λ(x)|1 for xB. Thus, if x0V¯, then there is some λX* such that λ(x0)>1 and if xV¯ then |λ(x)|1. From this it follows that λK, and then because λ(x0)>1 it follows that

x0{xX:if λK then |λ(x)|1}.

We have shown that if x0V¯ then x0 is not an element of the polar of K, and therefore

V¯={xX:if λK then |λ(x)|1}.

By the Banach-Alaoglu theorem, K is a compact set in (X*,𝒪X(X*)), and one checks that K is convex. The uniform boundedness principle1717 17 Walter Rudin, Functional Analysis, second ed., p. 46, Theorem 2.9 states that if

  • W and Z are topological vector spaces

  • K is a compact convex set in W

  • Γ is a set of continuous linear maps WZ

  • For each wK, the set Γ(w)={g(w):gΓ} is a bounded subset of Z

then there is a bounded set B in Z such that if gΓ then g(K)B. For xE, define gx:X* by gx(λ)=λ(x), and define Γ={gx:xE}. Because E is weakly bounded, for all λX* there is some γ(λ) such that if xE then |λ(x)|γ(λ); this is by Lemma 8. Hence for all λX* there is some γ(λ) such that the set Γ(λ) is contained in the closed disc in with radius γ(λ). Thus Γ(λ) is bounded in . We apply the uniform boundedness principle using W=(X*,𝒪X(X*)) and Z=. We thus obtain that there is some 0γ< such that if gΓ then g(K) is contained in the closed disc of radius γ. That is, for all xE and λK, we have

|gx(λ)|γ,

i.e. for all xE and λK we have

|λ(x)|γ.

Therefore, if xE and λK, then

|λ(1γx)|=1γ|λ(x)|1,

which means that 1γxV¯. This is true for all xE, so 1γEV¯N, and hence EγN. Since N was an arbitrary neighborhood of 0 in (X,𝒪(X)), we have satisfied the definition of the set E being bounded in (X,𝒪(X)), completing the proof. ∎

The final result we state in this section gives a condition for the dual of a locally convex space to be weak-* separable.1818 18 Walter Rudin, Functional Analysis, second ed., p. 90, chapter 3, Exercise 28. We already stated in Theorem 4 that the dual with the weak-* topology of a topological vector space is itself a locally convex space.

Theorem 10.

If X is a separable metrizable locally convex space, then (X*,OX(X*)) is a separable locally convex space.

5 Normed spaces

Let X be a normed space with norm . The topology on X is the coarsest topology such that for each x0X, the map xx-x0 is continuous. A normed vector space is locally convex and is metrizable, with metric d(x,y)=x-y.

If X and Y are topological vector spaces and T:XY is a linear map, we say that T is bounded if E being bounded in X implies that T(E) is bounded in Y.1919 19 A statement close to it is proved in Walter Rudin, Functional Analysis, second ed., p. 24, Theorem 1.32.

Theorem 11.

Let X and Y be normed vector spaces and let T:XY be linear. The following three statements are equivalent:

  • T is continuous.

  • T is bounded.

  • There is some γ such that if xX then Txγx.

If X and Y are normed spaces and T:XY is a bounded linear map, the operator norm T of T is defined to be the infimum of those γ such that if xX then Txγx. T=supx1Tx. The set of bounded linear maps XY is denoted (X,Y). (X,Y) is a normed space, and if Y is a Banach space then (X,Y) is a Banach space.2020 20 Walter Rudin, Functional Analysis, second ed., p. 92, Theorem 4. X*=(X,), and since is a Banach space, with the operator norm X* is a Banach space.

The following theorem rewrites Theorem 9 in the terminology of norms.2121 21 Walter Rudin, Functional Analysis, second ed., p. 71.

Theorem 12 (Weak boundedness is equivalent to boundedness).

If X is a normed space and E is a subset of X, then

supxE|λ(x)|<

for all λX* if and only if there is some γ such that xE implies

xγ.
Proof.

Suppose that supxE|λ(x)|< holds for all λX*. By Lemma 8, this means that the set E is weakly bounded. But by Theorem 9 this implies that E is bounded. The closed unit ball is a neighborhood of 0 in (X,𝒪(X)), so, as E is bounded, there is some γ such that E{γx:x1}. Hence, if xE then xγ.

Suppose that there is some γ such that xE implies that xγ, and let λX*. Because λ is continuous, for xE we have

|λ(x)|λxλγ<.

Theorem 13 (Banach-Alaoglu theorem).

If X be a normed vector space, then B={λX*:λ1} is a compact subset of (X,OX(X*)).

Proof.

It is a fact2222 22 Walter Rudin, Functional Analysis, second ed., p. 94, Theorem 4.3. (proved using the Hahn-Banach extension theorem) that for every xX,

x=sup{|λ(x)|:λB}.

From this and Theorem 5, it follows that B is weak-* compact in X*. ∎

The Eberlein-Smulian theorem states that a set in a normed space is weakly compact if and only if the set is weakly sequentially compact.2323 23 Robert E. Megginson, An Introduction to Banach Space Theory, p. 248, Theorem 2.8.6.

Theorem 14 (Eberlein-Smulian theorem).

If X is a normed space, then a subset A of X is compact in Ow(X) if and only if every sequence in A has a subsequence that converges in Ow(X) to an element of A.

6 Banach spaces

We say that a Banach space X is reflexive if X**=X, where X*=(X,) and X**=(X*,). (Although it makes sense to talk about X* for a normed space, X* is itself a Banach space and so too is X**, hence if a normed space were reflexive then it would have to be a Banach space.)

Kakutani’s theorem relates the property of a Banach space being reflexive with weak compactness.2424 24 Joseph Diestel, Sequences and Series in Banach Spaces, p. 18, chapter III.

Theorem 15 (Kakutani’s theorem).

The closed unit ball in a Banach space is weakly compact if and only if the space is reflexive.

Thus, combining the Eberlein-Smulian theorem and Kakutani’s theorem we get that a Banach space is reflexive if and only if the closed unit ball is weakly sequentially compact.

The following theorem states that a Banach space is separable if and only if the closed unit ball in the dual space is weak-* metrizable.2525 25 John B. Conway, A Course in Functional Analysis, second ed., p. 134, Theorem 5.1.

Theorem 16.

Let X be a Banach space and let B={λX*:λ1}. X is separable if and only if B with the subspace topology inherited from (X*,OX(X*)) is metrizable.

7 Hilbert spaces

Let H be a Hilbert space with inner product ,:H×H, linear in the first argument. The Riesz representation theorem2626 26 John B. Conway, A Course in Functional Analysis, second ed., p. 13, Theorem 3.4. states that if λH*, then there is a unique hλH such that if hH then

λ(h)=h,hλ,

and λ=hλ. On the other hand, if h0H then λ(h)=h,h0 satisfies, by the Cauchy-Schwarz inequality,

|λh0(h)|=|h,h0|hh0,

hence, by Theorem 11, λh0H*. Thus H*={λh0:h0H}. The weak topology 𝒪w(H) on H is the coarsest topology on H such that each λH* is continuous (H,𝒪w(H)), hence such that for each h0H, the funtion hh,h0 is continuous (H,𝒪w(H)). Thus, a net hαH converges to hH in 𝒪w(H) if and only if for all h0H we have that hα,h0 converges to h,h0, and this characterizes the weak topology on H.