Topological Vector Space - Note
Topological Vector Space - Note
PRADIPTA BANDYOPADHYAY
Definition 1.1. A topological vector space (tvs for short) is a linear space X
(over K) together with a topology J on X such that the maps (x, y) → x+y
and (α, x) → αx are continuous from X × X → X and K × X → X
respectively, K having the usual Euclidean topology.
Definition 1.3. A base for the topology at 0 is called a local base for the
topology J .
These notes are built upon an old set of notes by Professor AK Roy. The debt is
gratefully acknowledged.
1
2 BANDYOPADHYAY
Proof. Exercise.
Proof. We check that all the conditions of Theorem 1.4 are satisfied.
(i). U is clearly closed under finite intersections.
(ii). Suppose U = {x : pi (x) < δi , i = 1, . . . , n}. Let δ = mini δi /2 and
V = {x : pi (x) < δ, i = 1, . . . , n}. If y, z ∈ V then pi (y +z) ≤ pi (y)+pi (z) <
δ/2 + δ/2 = δ < δi ⇒ V + V ⊆ U .
(iii). Since pi (αx) = |α|pi (x), each U ∈ U is balanced.
(iv). If x ∈ X, pi (αx) = |α|pi (x) < δi if α is sufficiently small.
4 BANDYOPADHYAY
Example 1.10. (a) Let X be the vector space of all K-valued continuous
functions on a topological space Ω. For each compact set K ⊆ Ω, define
pK (f ) = supt∈K |f (t)|. The family {pK : K ⊆ Ω compact} gives a Hausdorff
vector topology in which convergence means uniform convergence on all
compact subsets of Ω. If K is restricted to finite subsets of Ω, we get the
topology of pointwise convergence. In general, if the sets K are restricted
to a class C of subsets of Ω, we obtain the topology of uniform convergence
on sets in C.
(b) Let X = C ∞ [a, b], the vector space of all infinitely differentiable (K-
valued) functions on the closed bounded interval [a, b]. For each n, define
pn (f ) = sup{|f (n) (t)| : t ∈ [a, b]} where f (n) is the n-th derivative of f . In
the topology defined by the pn , convergence means uniform convergence of
all derivatives.
We will now prove the converse of Theorem 1.8, that is, locally convex vector
topologies are generated by families of seminorms. But we first examine
convex sets in some detail.
Proof. (a). Let ε > 0 be given. There exists r > 0, s > 0 such that
r < pK (x) + ε/2, s < pK (y) + ε/2 and x/r, y/s ∈ K. Now,
x+y r x s y
= + ∈K
r+s r+s r r+s s
by convexity, hence pK (x + y) ≤ (r + s) < pK (x) + pK (y) + ε.
(b). If pK (x) < 1 then x/r ∈ K for some r < 1 ⇒ x = r(x/r) + (1 − r)(0) ∈
K. If y ∈ X then pK (x + λy) ≤ pK (x) + λpK (y) < 1 if λ > 0 is sufficiently
small, hence {pK < 1} ⊆ {x ∈ K : K is radial at x}. Conversely, if K is
radial at x, then x + λx ∈ K for some λ > 0, hence pK (x + λx) ≤ 1 by
1
definition of pK ⇒ pK (x) ≤ 1+λ < 1. By definition of pK , K ⊆ {pK ≤ 1}.
a x
(c). If x/r ∈ K and a 6= 0 ∈ K, then |a| r ∈ K (as K is balanced) ⇒
pK (ax) ≤ |a|r. Thus, pK (ax) ≤ |a|pK (x) (taking infimum over r).
Taking x/a instead of x, we get pK (x) ≤ |a|pK (x/a). Putting b = 1/a, we
get pK (bx) ≥ |b|pK (x).
(d). 0 ∈ K ◦ ⇒ there exists neighborhood U of 0 with U ⊆ K. Let ε > 0
be given. If y ∈ εU then pK (y) = pK (εu) = εpK (u) ≤ ε (since x ∈ K ⇒
pK (x) ≤ 1) ⇒ pK is continuous at 0 ⇒ pK continuous everywhere.
pK continuous ⇒ {pK ≤ 1} is closed ⇒ K ⊆ {pK ≤ 1}. Suppose pK (x) ≤ 1.
If 0 < λ < 1, then pK (λx) = λpK (x) < 1 ⇒ λx ∈ K. If λ → 1 then λx → x,
hence x ∈ K ⇒ K ⊇ {pK ≤ 1}, hence K = {pK ≤ 1}.
pK continuous ⇒ {pK < 1} is open and hence ⊆ K ◦ . But if pK (x) = 1 then
xn = x/(1 − 1/n) ∈ / K as pK (xn ) = 1/(1 − 1/n) > 1 and K ⊆ {pK ≤ 1},
but xn → x, so x is a limit of points not in K, hence x 6∈ K ◦ .
6 BANDYOPADHYAY
Proof. If X has a local base consisting of convex sets, it has a local base B
consisting of closed convex balanced neighborhoods of 0. For U ∈ B, the
Minkowski functional pU is a seminorm. Since U = {pu ≤ 1}, the family
{pU : U ∈ B} generates the topolgy of X.
The only trouble is that the balls {x : d(0, x) ≤ r} need not be convex (they
are balanced though) as we see from the following example:
Example 1.18. Let s = {(xn )∞ n=1 : xn ∈ K for all n ≥ 1}, the space of all
scalar sequences. The topology of pointwise convergence is described by the
seminorms pk , (k ≥ 1), pk ((xn )) = |xk | and the metric is
X 1 |xn − yn |
d(x, y) = , x = (xn ), y = (yn ).
2n 1 + |xn − yn |
TOPOLOGICAL VECTOR SPACES 7
Remark 1.20. The notion of a Cauchy net for a tvs (X, J ) can be defined
without reference to any metric. Fix a local base F at 0 for the topology J .
A net {xα } is said to be (J -) Cauchy if, for any U ∈ F, there exists α0 such
that α ≥ α0 and β ≥ α0 ⇒ xα − xβ ∈ U . It is clear that different local bases
for J give rise to the same class of Cauchy nets. Now let (X, J ) be metrized
by an invariant metric d. As d is invariant and the d-balls centered at 0 from
a local base, we conclude that a sequence {xn } ⊆ X is a d-Cauchy sequence
if and only if it is a J -Cauchy sequence. Consequently any two invariant
metrics on X that are compatible with the topology have the same Cauchy
sequences and the same convergent sequences, viz. the J -convergent ones.
By far the most widely discussed locally convex spaces are those for which
the vector topologies are given by a single norm, the so-called normed spaces.
Proof. Let X be normed by k · k. Then the open unit ball {x : kxk < 1} is
a convex and bounded neighborhood of 0.
For the converse, let V be a bounded convex open neighborhood of 0. By
Lemma 1.5 (xiii), there exists neighborhood U of 0 such that U ⊆ V , U
convex, open and balanced. Obviously, U is also bounded. For x ∈ X,
define kxk = pU (x) where pU is the Minkowski functional of U .
Claim: {λU : λ > 0} form a local base for the topology J of X.
8 BANDYOPADHYAY
Exercise 1. Show that s, the space of all scalar sequences, is not normable
where the topology on s is defined by the metric
∞
X 1 |xn |
d(0, {xn }) = .
2n 1 + |xn |
n=1
2. Quotient Spaces
Lemma 2.1. X/M is a tvs, π is a continuous and open map, and X/M is
Hausdorff if and only if M is closed.
TOPOLOGICAL VECTOR SPACES 9
¯
d([x], [y]) = d(x + M, y + M ) = d(x − y, M )
Inductively choose xi ∈ X so that d(xi , xi+1 ) < 2−i and π(xi ) = [uni ]. Since
d is complete, xi → x ∈ X and as π is continuous, π(xi ) = [uni ] → π(x).
Hence {[un ]}, being Cauchy, must converge to π(x).
3. Duals of tvs
Proof. (i) ⇒ (ii) and (iv) ⇒ (v) trivial. (ii) ⇒ (iii) clear as f 6≡ 0.
(iii) ⇒ (iv). Choose x ∈ X and a balanced neighborhood U of 0 such that
(x + U ) ∩ ker(f ) = ∅ ⇒ f (x) ∈
/ −f (U ) (which proves (v)). But f (U ) is
balanced as U is balanced and hence f (U ) is bounded [as a proper balanced
subset of K must be bounded] which proves (iv).
(v) ⇒ (i). Assume that f maps a balanced neighborhood U of 0 onto a
proper subset of K. Hence f (U ) is bounded. We have a k > 0 such that
f (U ) ⊆ {z ∈ K : |z| ≤ k}. Let ε > 0. Then f ( kε U ) ⊆ {z ∈ K : |z| ≤ ε} ⇒ f
is continuous at 0.
(vi) ⇔ (i) follows from the observation that f (x) = Ref (x) − iRef (ix).
Remark 3.2. The last observation that f (x) = Ref (x)−iRef (ix) is useful.
We can consider a complex vector space X as a vector space over the R
by restricting scalar multiplication to R and the space XR thus obtained
is called the real restriction of X. The above shows that (X ∗ )R is (real)
linearly isomorphic to (XR )∗ under the map f → Ref .
TOPOLOGICAL VECTOR SPACES 11
We want to know conditions which ensure that there exists non-zero contin-
uous linear functionals on a tvs.
Remark 3.4. This result makes it clear why lctvs are important viz. these
have plenty of non-zero continuous linear functionals.
Exercise 5. When 0 < p < 1, Lp (dµ) is defined in the usual way but because
of the failure of Minkowski’s inequality for such values of p, one does not
get a norm. However, d(f, g) = |f − g|p dµ does define an invariant metric
R
on Lp (dµ). Show by using (ii) of Theorem 3.3, that (Lp [0, 1])∗ = {0}, with
µ as Lebesgue measure, in contrast to the case when 1 ≤ p < ∞.
Solution: Let Uε = {f : |f |p dµ < ε} is a basic neighborhood. Enough to
show that co(Uε ) is the whole space. Let f ∈ Lp (dµ) and let |f |p dµ = M .
R
Rx p
Since x 0 |f (y)| Rdµ(y) is a continuous function, there exists a point
x
x0 ∈ [0, 1] such that 0 0 |f |p dµ = M 2 . Subdivide [0, x0 ] and [x0 , 1] further
n
in this way to obtain 2 intervals In (n to be determined) in each of which
R p M P2n n
In |f | dµ = 2n . Let gi = f χIn . Then f = i=1 gi . Put hi = 2 gi . Then
2n Z Z
X 1 p M
f= n
hi and |hi | dµ = 2np |fi |p dµ = n(1−p) < ε
2 Ii 2
i=1
Thus fe ≤ p ⇐⇒
Now, if u, v ∈ Y ,
5. Consequences
g|M = f, |g(x)| ≤ 1 ∀ x ∈ A.
Exercise 8. A closed convex set in a lctvs is the intersection of all the closed
half spaces containing it.
S∞
Proof. A radial at 0 and A balanced ⇒ X = n=1 nA.
A is balanced (as Ti is linear for each i), closed (as each Ti is continuous
and U is closed) and radial at 0 because if x ∈ X, there exists rx > 0
such that rx (Ti x) ∈ U (for all i) by the boundedness of {Ti x : i ∈ I}
in Y . So by last Proposition, A + A is a neighborhood of 0 and clearly
A + A ⊆ i∈I Ti−1 (V ).
T
Remark 6.4. The above result is often called the Principle of Uniform
Boundedness when X, Y are, respectively, Banach and normed linear spaces.
Specifically, in this context, the theorem reads:
If {Ti }i∈I is a family of continuous linear transformations from X to Y such
that for each x ∈ X, sup{kTi xkY : i ∈ I} < ∞ then there exists M > 0 such
that kTi xkY ≤ M kxkX (for all i and x ∈ X), hence supi kTi k < ∞.
Exercise 10. Suppose Y is complete metrizable tvs such that the contin-
uous functionals on Y separate points of Y , i.e. given y1 6= y2 there exists
f ∈ Y ∗ such that f (y1 ) 6= f (y2 ). If X is a complete tvs then T : X → Y ,
(T assume to be linear) is continuous if and only if f ◦ T is continuous for
all f ∈ Y ∗ .
TOPOLOGICAL VECTOR SPACES 19
P
Exercise 11. Prove that @ a sequence {λn } ⊆ C such that an converges
absolutely if and only if {λn an } is bounded.
[Hint: Assume (λn ) exists with λn 6= 0 for all n. Define T : `∞ → `1 by
T [(cn )] = {cn /λn }. Then kT k ≤ | λ1n | and
P P 1
| λn | < ∞ by the hypothesis.
Hence T is continuous, linear, one-one, onto ⇒ `∞ and `1 are homeomor-
phic which is impossible as `∞ is non-separable and `1 is separable in their
respective norm topologies.]
α1 , . . . , αn ∈ K.
Exercise 13. Let X be a lctvs and E ⊆ X is a convex set. Then the closure
of E in the weak and the original topology are the same.
Definition 7.4. Suppose that X is a lctvs For each x ∈ X, define the linear
functional x̂ on X ∗ by x̂(x∗ ) = x∗ (x). Note that {x̂ : x ∈ X} is a linear space
and that it separates the points of X ∗ . We are now in the situation described
by the lemma with X replaced by X ∗ and F replaced by {x̂ : x ∈ X}. The
corresponding topology is called the weak* topology on X ∗ . A basic w*-
neighborhood of 0 is then {x∗ : |x∗ (xk )| < ε, xk ∈ X, k = 1, . . . , n}.
TOPOLOGICAL VECTOR SPACES 21
Corollary 7.6. Let X be a normed linear space. Then the unit ball of X ∗
is weak* compact.
Exercise 14. If X is a separable normed linear space. Show that the w*-
topology of the unit ball of X ∗ is metrizable.
Example 8.2. Let K be a compact Hausdorff space and let P be the convex
set of regular probability measures on K, so P ⊆ X where X = C(K)∗ .
(i) The discrete measures form a face of P (a measure µ in P is
discrete if µ = ∞
P P
n=1 αn δxn , n αn = 1, αn ≥ 0, xn ∈ K).
(ii) The continuous measures (i.e., measures without any point
masses) from a face of P .
(iii) If m ∈ P , then {µ ∈ P : µ m} is a face of P .
(iv) If m ∈ P , then {µ ∈ P : µ ⊥ m} is a face of P .
(v) The extreme points of P are the point masses {δx : x ∈ K} and
conversely.
22 BANDYOPADHYAY
(vi) Using (v), show that the extreme points of the unit ball of
C(K)∗ = M (K) are exactly {λδx : λ ∈ C, |λ| = 1, x ∈ K} = B.
(vii) Consider a subspace A of C(K). Show that the extreme points
of the unit ball of A∗ are contained in B and find an example to
show that all points in B may not be extreme.
(The proof of (vii) follows easily from the Krein-Milman theorem)
9. Integral Representations
Proof. Suppose µ ∈ M1+ (C). Then by the last result, r(µ) = x ∈ co(C).
Now, let x ∈ co(C). Then x is approximable by elements of the form
yα = ni=1
P α α α α P α
λi yi (αi ≥ 0, λi = 1, yiα ∈ C). Consider the corresponding
λα δyiα ∈ M1+ (C), r(µα ) = yα . There exists a subnet
P i
measures µα =
µβ → µ ∈ M1+ (C), hence for all f ∈ X ∗ , limβ f (yβ ) = limβ µβ (f ) = µ(f ).
But yβ → x (as yα → x), and we have µ(f ) = f (x) for all f ∈ X ∗ .
Remark 9.3. It now follows from the above result that the following state-
ments are equivalent:
(a) If K ⊆ X is a compact convex set, then K = co(ext(K)).
(b) Each x ∈ K is the resultant of a µ ∈ M1+ (K) with µ supported
by ext(K).