notes-B-2025-02-10
notes-B-2025-02-10
T
is a metric on X . The resulting topology is the norm topology and it is the default
topology on X .
AF
Definition 1.2. A normed vector space X is a Banach space if it is complete (with its
norm topology). □
for all x ∈ X . □
Remark 1.4. Equivalent norms determine the same topology on X and the same
Cauchy sequences (Problem 1.4). In particular, it follows that if X is equipped with
two equivalent norms ∥ · ∥1 , ∥ · ∥2 then it is complete (a Banach space) in one norm if
and only if it is complete in the other.
Equivalence of norms is an equivalence relation on the set of norms on X . □
The next proposition is simple but fundamental; it says that the norm and the
vector space operations are continuous in the norm topology.
Proposition 1.5 (Continuity of vector space operations). Let X be a normed vector
space over F.
a) If (xn ) converges to x in X , then (∥xn ∥) converges to ∥x∥ in R.
b) If (kn ) converges to k in F and (xn ) converges to x in X , then (kn xn ) converges
to kx in X .
c) If (xn ) converges to x and (yn ) converges to y in X , then (xn + yn ) converges to
x + y in X .
1
2
Proof. The proofs follow readily from the properties of the norm, and are left as exercises.
□
The following proposition gives a convenient criterion for a normed vector space to
be complete.
Definition 1.6. Given a sequence (xn ) from a normed vector space X , the expression
P PN
n=1∞ xn denotes the sequence (sN = n=1 xn ), called the sequence of partial sums of
the series. The series converges Pif the sequence of partial sums converges in the norm
topology. In this case we use n=1∞ xn to also denote the limit of this sequence and
call it the sum.
Explicitly, the series ∞
P
n=1 xn converges means there is an x ∈ X such that for each
ϵ > 0 there is an N such that ∥sn − x∥ < ϵ for all n ≥ N.
The series ∞
P P∞
n=1 xn converges absolutely if the series n=1 ∥xn ∥ converges (in the
normed vector space (R, | · |)). □
Before proving the Proposition we collect two lemmas. A definition is needed for
the first.
Lemma 1.9. If (xn ) is a Cauchy sequence from a normed vector space X , then there is
a subsequence (yk ) of (xn ) that is super-cauchy.
Proof. With ϵ = 21 , there exists an N1 such that ∥xn − xm ∥ < 12 for all m, n ≥ N1 since
(xn ) is Cauchy. Assuming N1 < N2 < · · · < Nk have been chosen so that ∥xn − xm ∥ < 21j
1
for 1 ≤ j ≤ k, there is an Nk+1 < Nk such that ∥xn − xm ∥ < 2k+1 since (xn ) is Cauchy.
Hence by recursion we have constructed a (strictly)increasing sequence of integers Nk
such that ∥xn −xm ∥ < 21k for all m, n ≥ Nk . Set yk = xNk and note that ∥yk+1 −yk ∥ < 21k ,
from which it follows that (yk ) is a super-cauchy subsequence of (xn ). □
The proof will also use the following standard lemma from advanced calculus.
Lemma 1.10. If (xn ) is a Cauchy sequence from a metric space (X, d) and if (xn ) has
a subsequence (yk ) that converges to some x, then (xn ) converges to x.
then
N
X N
X
∥sN − sM ∥ = xn ≤ ∥xn ∥ < ϵ.
n=M +1 n=M +1
Thus the sequence (sN ) is Cauchy in X , hence convergent by the completeness hypoth-
esis.
Conversely, suppose every absolutely convergent series in X is convergent and that
(xn ) is given Cauchy sequence from X. By Lemma 1.9 there is a super-cauchy subse-
quence (yk ) of (xn ). Since (yk ) is super-cauchy, the series ∞
P
k=1 (yk+1 − yk ) is absolutely
convergent and hence, by hypothesis, convergent in X . Thus there is an z ∈ X such
that the sequence of partial sums
X n
(yk+1 − yk ) = yn+1 − y1
k=1
1.2. Examples.
1.2.1. Euclidean space. Observe that the Euclidean norm on the complex vector space
Cn agrees with the Euclidean norm on the real vector space R2n (via that natural real
linear isomomorphism R2 → C sending (x, y) to x + iy). Thus, Fn with the usual
1/2
Euclidean norm ∥(x1 , . . . xn )∥ = ( nk=1 |xk |2 ) is a Banach space.
P
Given a normed vector space X = (X , ∥ · ∥), denote its closed unit ball by
X1 = {x ∈ X : ∥x∥ ≤ 1}.
4
It is instructive to sketch the closed unit ball in R2 with the three norms above.
It turns out that any two norms on a finite-dimensional vector space are equivalent.
As a corollary, every finite-dimensional normed space is a Banach space. See Problem 1.5.
Lemma 1.11. If ∥·∥1 and ∥·∥2 are norms on X and there is a constant C > 0 such that
∥x∥1 ≤ C∥x∥2 for all x ∈ X , then the mapping ι : (X , ∥ · ∥2 ) → (X , ∥ · ∥1 ) is (uniformly)
continuous.
Proof. For x, y ∈ X , we have ∥ι(x) − ι(y)∥1 = ∥ι(x − y)∥1 = ∥x − y∥1 ≤ C∥x − y∥2 . □
Proposition 1.12. If ∥x∥ is a norm on Rn , then ∥x∥ is equivalent to the Euclidean
norm ∥ · ∥2 .
Sketch of proof. Let {e1 , . . . , en } denote the usual basis for Rn . Given x = aj ej ∈ Rn ,
P
X X
∥x∥ ≤ |aj | ∥ej ∥ = |aj |∥ej ∥ ≤ M ∥x∥1 ≤ n M ∥x∥2 ,
where M = max{∥e1 ∥, . . . , ∥en ∥}. It now follows that the map ι : (Rn , ∥ · ∥2 ) → (Rn , ∥ · ∥)
is continuous and therefore so is the map f : (Rn , ∥ · ∥2 ) → [0, ∞) defined by f (x) = ∥x∥.
Since
S n−1 = {x ∈ Rn : ∥x∥2 = 1}
(the unit sphere) is compact in Rn , by the Extreme Value Theorem, f attains its infimum;
that is, there is a point p ∈ S n−1 such that f (p) ≤ f (x) for all x ∈ S n−1 . But f (p) =
∥p∥ > 0 since p ̸= 0. Let c = f (p) = ∥p∥. We conclude that if ∥x∥2 = 1 then ∥x∥ ≥ c∥x∥2 ,
from which it follows by homogeneity that ∥x∥ ≥ c∥x∥2 for all x ∈ Rn . □
Corollary 1.13. All norms on a finite dimensional vector space are equivalent. Further,
if V is a finite dimensional normed vector space, then V1 is compact and V is a Banach
space.
Proof. Suppose V is a normed vector space of dimension n and let {v1 , . . . , vn } denote
a basis for V . The function ∥ · ∥′ : V → [0, ∞) defined by
X X
∥v∥′ = ∥ aj vj ∥′ = |aj |
is easily seen to be a norm.
Now let ∥ · ∥ be a given norm on V. This norm induces a norm ∥ · ∥∗ on Rn given by
X X
∥ aj ej ∥∗ = ∥ aj vj ∥.
Since all norms in Rn are equivalent, the norm ∥ · ∥∗ is equivalent to the norm ∥ · ∥1 .
Hence there exist constants 0 < c < C such that
X X X X X
c∥v∥′ = c |aj | = c∥ aj ej ∥1 ≤ ∥ aj ej ∥∗ ≤ C∥ aj ej ∥1 = C |aj | = C∥v∥′ .
P P
Thus, as ∥ aj ej ∥∗ = ∥ aj vj ∥,
c∥v∥′ ≤ ∥v∥ ≤ C∥v∥′
5
1.2.2. The Banach space of bounded functions. If V is a vector space over F and ∅ ̸= T
is a set, then F (T, V ), the set of functions f : T → V is a vector space over F under
pointwise operations; e.g., if f, g ∈ F (T, V ) then f + g : T → V, is the function defined
by (f + g)(t) = f (t) + g(t).
Definition 1.14. A subset R of a normed vector space X is bounded if there is a C
such that ∥x∥ ≤ C for all x ∈ R; that is, R ⊆ CX1 .
A function f : T → X is bounded if f (T ) ⊆ X is bounded.
Let Fb (T, X ) denote the vector space (subspace of F (T, X )) of bounded functions
f : T → X.
Remark 1.15. The function ∥ · ∥∞ : Fb (T, X ) → [0, ∞) defined by
∥f ∥∞ = sup{|f (t)| : t ∈ T }
is a norm on Fb (T, X ) as you should verify. Let d∞ denote the resulting metric:
d∞ (f, g) = ∥f − g∥∞ .
Note that convergence of a sequence in the metric space (Fb (T, X ), d∞ ) is uniform
convergence; in particular, a sequence is Cauchy in Fb (T, X ) if and only if it is uniformly
Cauchy. (Exercise.)
Proposition 1.16. If X is a Banach space, then Fb (T, X ) is also Banach space.
Given s ∈ T, there is an M ≥ N such that ∥fm (s) − f (s)∥ < ϵ for all m ≥ N. Since,
(fm (s) − fn (s))m converges (with m) to (f (s) − fn (s)) in X , another application of
Proposition 1.5 gives (∥fm (s) − fn (s)∥)m converges to ∥f (s) − fn (s)∥. Thus
∥f (s) − fn (s)∥ ≤ ϵ,
for all s ∈ T. Hence d∞ (f, fn ) = ∥f − fn ∥ ≤ ϵ and the proof is complete. □
There are important Banach spaces of continuous functions. Before going further,
we remind the reader of the following result from advanced calculus.
Theorem 1.17. Suppose X, Y are metric spaces, (fn ) is a sequence fn : X → Y and
x ∈ X. If each fn is continuous at x and if (fn ) converges uniformly to f , then f is
continuous at x. Hence if each fn is continuous, then so is f.
Proof. Let x and ϵ > 0 be given. Choose N such that if n ≥ N and y ∈ X, then
dY (fn (y), f (y)) < ϵ. Since fN is continuous at x, there is a δ > 0 such that if dX (x, y) < δ,
then dY (fN (x), fN (y)) < ϵ. Thus, if dX (x, y) < δ, then
dY (f (x), f (y)) ≤ dY (f (x), fN (x)) + dY (fN (x), fN (y)) + dY (fN (y), f (y))
< 3ϵ,
proving the theorem. □
Given a normed vector space Y, let Cb (X, Y) denote the subspace of Fb (X, Y) con-
sisting of continuous functions. Since uniform convergence is the same as convergence in
the normed vector space (Fb (X, Y), d∞ ), by Theorem 1.17, Cb (X, Y) is a closed subspace
of Fb (X, Y). In particular, in the case Y is a Banach space, so is Cb (X, Y).
When X be a compact metric space, let C(X) = C(X, F) denote the set of con-
tinuous functions f : X → F. Thus C(X) is a subspace of Fb (X, F) and we endow
C(X) with the norm it inherits from Fb (X, F). Since F is complete, C(X) is a Banach
space. Of course here we could replace F by a Banach space X and obtain the analogous
conclusion for the space C(X, X ).
Now let X be a locally compact metric space. In this case, a function f : X → F
vanishes at infinity if for every ϵ > 0, there exists a compact set K ⊆ X such that
/ |f (x)| < ϵ. Let C0 (X) denote the subspace of Fb (X, F) consisting of continuous
supx∈K
functions f : X → F that vanish at infinity. Then C0 (X) is a normed vector space with
the norm it inherits from C(X) (equivalently Fb (X, F). It is routine to check that C0 (X)
is complete.
1.2.3. L1 spaces over R. Let (X, M , µ) be a measure space and let L1 (µ) denote the
(real) vector space of (real-valued) absolutely integrable functions on X from Theo-
rem ??. We saw that Z
∥f ∥1 := |f | dm
X
7
defines a norm on L1 (µ), provided we agree to identify f and g when f = g a.e. (Indeed
the chief motivation for making this identification is that it makes ∥ · ∥1 into a norm.
Proposition 1.18. The real vector space L1 (µ) is a Banach space.
1.2.4. Complex L1 (µ) spaces. In this subsection we describe the extension of L1 (µ) to a
complex vector space of complex valued functions (equivalence classes of functions).
Again we work on a fixed measure space (X, M , µ). As a topological space, C and
R , are the same. A function f : X → C = R2 is measurable if and only if it is M − B2
2
measurable. Measurability of f can also be described in terms of the real and imaginary
parts of f.
Proposition 1.19. Suppose (X, M ) is a measurable space and f : X → C. Writing
f : X → C as f = u + iv, where u, v : X → R, the function f is measurable if and only
if both u and v are.
Moreover, if f is measurable, then so is |f | : X → [0, ∞).
We begin with the following elementary lemma whose proof is left to the reader.
Lemma 1.20. Suppose (X, M ) is a measure space and Y and Z are topological spaces.
If f : X → Y is M − BY measurable and g : Y → Z is BY − BZ measurable, then g ◦ f
is M − BZ measurable. In particulr, the result holds if g is continuous.
Sketch of proof of Proposition 1.19. The Borel σ-algebra B2 is generated by open rect-
angles; that is, a set U ⊆ C is open if and only if it is a countable union of open rectangles
(with rational vertices even). For an open rectagle I = J × K = (a, b) × (c, d) observe
that
f −1 (J) = u−1 (J) ∩ v −1 (K).
Thus, if u and v are measurable, then f −1 (J) ∈ M . Consequently, by Propition ??, f
is measurable. Hence if u, v are both measurable, then so is f.
8
Likewise Λ(f + g) = Λ(f ) + Λ(g). Thus Λ is C-linear on L1C and item (a) is proved.
R
If R f = 0, then f = 0 almost everywhere and the last three items hold. Otherwise,
write f = reit in polar coordinates and observe
Z
−it
e f ∈ R+ .
Hence item (c) holds. Similarly,the triangle inequality, item (d), follows from |f + g| ≤
|f | + |g| (pointwise). □
Remark 1.25. Proposition 1.24 says ∥ · ∥1 is a semi-norm on L1 . As usual, we identify
functions that differ by a null vector; that is, f ∼ g if ∥f − g∥1 = 0; equivalently,
identifying functions that are equal a.e., we obtain a normed complex vector space of L1
functions (which of course are not actually functions).
Along with these spaces it is also helpful to consider the vector space
Notice that c00 is a vector subspace of each of c0 , ℓ1 and ℓ∞ . Thus it can be equipped with
either the ∥ · ∥∞ or ∥ · ∥1 norms. It is not complete in either of these norms, however.
What is true is that c00 is dense in c0 and ℓ1 (but not in ℓ∞ ). (See Problem 1.11).
1.2.6. Lp spaces. Again let (X, M , m) be a measure space. For 1 ≤ p < ∞ let Lp (m)
denote the set of measurable functions f for which
Z 1/p
p
∥f ∥p := |f | dm <∞
X
(again we identify f and g when f = g a.e.). It turns out that this quantity is a norm on
Lp (m), and Lp (m) is complete, though we will not prove this yet (it is not immediately
obvious that the triangle inequality holds when p > 1).
Choosing (X, M , µ) = (N, P (N), c), counting measure on N, obtains the sequence
spaces ℓp ; that is, the F-vector space of functions f : N → F such that
∞
!1/p
X
p
∥f ∥p := |f (n)| <∞
n=1
as for the other Lp spaces we identify f and g when there are equal a.e. When f ∈ L∞ ,
let ∥f ∥∞ be the smallest M for which (1) holds. Then ∥ · ∥∞ is a norm making L∞ (µ)
into a Banach space.
is a norm on X ⊕ Y. These three norms are equivalent; indeed it follows from the
definitions that
∥(x, y)∥∞ ≤ ∥(x, y)∥2 ≤ ∥(x, y)∥1 ≤ 2∥(x, y)∥∞ .
If X and Y are both complete, then X ⊕ Y is complete in each of these norms. The
resulting Banach spaces are denoted X ⊕∞ Y, X ⊕1 Y and X ⊕2 Y.
Since the three norms in the previous paragraph are equivalent, the resulting spaces
are indistinguishable topologically. There is a more abstract description of this topology.
Definition 1.26. Given topological spaces (X, τ ) and (Y, σ), the product topology on the
Cartesian product X × Y is the smallest topoology that makes the coordinate projections
πX : X × Y → X and πY : X × Y → Y defined by πX (x, y) = x, πY (x, y) = y continuous.
That is, the topology generated by the sets U × Y and X × V for open sets U ⊆ X and
V ⊆ Y.
Proposition 1.27. Suppose (X, τ ) and (Y, σ) are topological spaces. The collection of
sets
B = {U × V : U ⊆ X, V ⊆ Y are open}
is a base for the product topology.
For normed vector spaces X and Y, the product topology on X × Y is metrizable and
is the norm topology on X × Y with any of the norms of equation (2). Consequentely,
a sequence zn = (xn , yn ) from X × Y converges (in the product topology) if and only if
both (xn ) and (yn ) converge; and zn converges to z = (x, y) if and only if (xn ) converges
to x and (yn ) converges to y. In particular, if X and Y are Banach spaces, then so is
X × Y.
It is evident how to extend the discussion here to finite products. The product topol-
ogy is the default topology on (finite) products of Banach spaces (and more generally
normed vector spaces).
1.2.8. Qoutient spaces. If X is a normed vector space and M is a proper subspace, then
one can form the algebraic quotient X /M, defined as the collection of distinct cosets
{x + M : x ∈ X }. From linear algebra, X /M is a vector space under the standard
operations. Let π : X → X /M denote the quotient map.
12
inf ∥f − g∥ = 0
g∈M
Proof. We will verify a couple of the axioms of a norm for the quotient norm, leaving the
remainder of the proof as an exercise. First suppose x ∈ X and ∥π(x)∥ = 0. It follows
that there is a sequence (mn ) form M such that (∥x − mn ∥) converges to 0; that is, (mn )
converges to x. Since M is closed, x ∈ M and hence π(x) = 0.
Now let x, y ∈ X and ϵ > 0 be given. There exists m, n ∈ M such that
∥x − m∥ ≤ ∥π(x)∥ + ϵ, ∥y − n∥ ≤ ∥π(y)∥ + ϵ.
Hence
from which it follows that the triangle inequality holds and we have proved the quotient
norm is indeed a norm.
To prove X /M is complete (with the quotient norm) under the assumption that
P
X is a Banach space (complete), suppose (yn ) is a sequence from X /M and yn is
1
absolutely convergent. For each n there exists xn ∈ X such that ∥xn ∥ ≤ ∥yn ∥ + n2 and
P
π(xn ) = yn . It follows that xn is absolutely convergent. Since X is a Banach space
the sequence of partial sums sN = N
P
n=1 xn converges to some x ∈ X . In partiulcar,
N
X
∥sN − x∥ ≥ ∥π(sN − x)∥ = ∥π(sN ) − π(x)∥ = ∥ yn − π(x)∥.
n=1
P
Since (∥sN − x∥) converges to 0, it follows that yn converges to π(x). Hence X /M is
complete by Proposition 1.7. □
More examples are given in the exercises and further examples will appear after the
development of some theory.
13
Remark 1.30. Note that in Definition 1.29 it suffices to require that ∥T x∥Y ≤ C∥x∥X
just for all x ̸= 0, or for all x with ∥x∥X = 1 (why?). □
Proof. Suppose T is bounded; that is, there exists a C ≥ 0 such that ∥T x∥ ≤ C∥x∥ for
all x ∈ X . Thus, if x, y ∈ X , then, ∥T x − T y∥ = ∥T (x − y)∥ ≤ C∥x − y∥ by linearity of
T . Hence (i) implies (ii). The implications (ii) implies (iii) implies (iv) implies (v) are
evident.
The proof of (v) implies (i) exploits the homogeneity of the norm and the linearity
of T and not nearly the full strength of the continuity assumption. By hypothesis, with
ϵ = 1 there exists δ > 0 such that if ∥x∥ = ∥x − 0∥ < 2δ, then ∥T x∥ = ∥T x − T 0)∥ < 1.
Given a nonzero vector x ∈ X , the vector δx/∥x∥ has norm less than δ, so
δx ∥T x∥
1> T =δ .
∥x∥ ∥x∥
1The suprema need not be attained.
14
The set of all bounded linear operators from X to Y is denoted B(X , Y). It is a
vector space under the operations of pointwise addition and scalar multiplication. The
quantity ∥T ∥ is easily seen to be a norm. It is called the operator norm of T .
Problem 1.1. Prove the ∥ · ∥1 and ∥ · ∥∞ norms on c00 are not equivalent. Conclude
from your proof that the identity map on c00 is bounded from the ∥ · ∥1 norm to the
∥ · ∥∞ norm, but not the other way around.
Problem 1.2. Consider c0 and c00 equipped with the ∥ · ∥∞ norm. Prove there is no
bounded operator T : c0 → c00 such that T |c00 is the identity map. (Thus the conclusion
of Proposition 1.33 can fail if Y is not complete.)
Proposition 1.32. For normed vector spaces X and Y, the operator norm makes
B(X , Y) into a normed vector space that is complete if Y is complete.
Proof. That B(X , Y) is a normed vector space follows readily from the definitions and
is left as an exercise.
Suppose now Y is complete, and let Tn be a cauchy sequence in B(X , Y). Let
E = X1 denote the closed unit ball in X . For x ∈ E,
(4) ∥Tn x − Tm x∥ = ∥(Tn − Tm )x∥ ≤ ∥Tn − Tm ∥∥x∥ ≤ ∥Tn − Tm ∥.
Hence the sequence (Tn |E ) is a cauchy sequence (so uniformly cauchy) from Cb (E, Y),
the space of bounded continuous functions from B to Y. Since Y is complete, there is
an F ∈ Cb (E, Y) such that (Tn |E ) converges to F in Cb (E, Y) and moreover ∥F (x)∥ ≤
C := sup{∥Tn ∥ : n} < ∞. See Subsection 1.2.2. An exercise shows, given x, y ∈ B and
c ∈ F if x + y ∈ B and c x ∈ B, then F (x + y) = F (x) + F (y) and F (cx) = cF (x).
Hence F extends, by homogeneity, to a linear map T : X → Y such that ∥T ∥ ≤ C and,
by equation (4), (Tn ) converges to T in B(X, Y ). □
If T ∈ B(X , Y) and S ∈ B(Y, Z), then two applications of the inequality (3) give,
for x ∈ X ,
∥ST x∥ ≤ ∥S∥∥T x∥ ≤ ∥S∥∥T ∥∥x∥
15
and it follows that ST ∈ B(X , Z) and ∥ST ∥ ≤ ∥S∥∥T ∥. In the special case that Y = X
is complete, B(X ) := B(X , X ) is an example of a Banach algebra.
The following proposition is very useful in constructing bounded operators—at least
when the codomain is complete. Namely, it suffices to define the operator (and show
that it is bounded) on a dense subspace.
Proposition 1.33 (Extending bounded operators). Let X , Y be normed vector spaces
with Y complete, and E ⊆ X a dense linear subspace. If T : E → Y is a bounded linear
operator, then there exists a unique bounded linear operator Te : X → Y extending T (so
Te|E = T ). Further ∥Te∥ = ∥T ∥.
surjective. As examples, let ℓp = ℓp (N) denote the sequence spaces from Subsection 1.2.5.
The linear map S : ℓp (N) → ℓp (N) defined by Sf (n) = 0 if n = 0 and f (n − 1) if n > 0
(for f = (f (n))n ∈ ℓp ) is the shift operator . It is straightforward to verify that S is an
isometry but not onto.
Example 1.39. Following up on the previous example, a linear map T : X → X can
be one-one and have dense range without being (boundedly) invertible. Let en ∈ ℓ2 (N)
denote the function en (m) = 1 if n = m and 0 otherwise for non-negative integers
0 ≤ m, n. The set of c00 = { N 2
P
n=0 an en : N ∈ N, cn ∈ F} is dense in ℓ (N) and the
mapping D : c00 → ℓ2 (N) defined by
N N
X X an
D( an e n ) = en
n=0 n=0
n+1
is easily seen to be bounded with ∥D∥ = 1. It is also injective. Hence D extends to an
injective bounded operator, still denoted D, from ℓ2 → ℓ2 , with ∥D∥ = 1. The range of
D contains {en : n ∈ N} and is thus dense in ℓ2 (N).
Since ∞
P 1 2
P∞ 1 2
n=0 | n+1 | < ∞, the vector f = n=1 n+1 en is in ℓ (N). On the other hand,
if g ∈ ℓ2 (N) and Dg = f, then
1 1
g(n) = (Dg)(n) = f (n) =
n+1 n+1
2
P
and thus g(n) = 1 for all n; however, since |g(n)| = ∞, we obtain a contradiction.
Hence f is not in the range of D.
1.4. Examples.
(a) If X is a finite-dimensional normed space and Y is any normed space, then every
linear transformation T : X → Y is bounded. See Problem 1.16.
(b) Let X denote c00 equipped with the ∥ · ∥1 norm, and Y denote c00 equipped with the
∥ · ∥∞ norm. Then the identity map idX ,Y : X → Y is bounded (in fact its norm is
equal to 1), but its inverse, the identity map ιY,X : Y → X , is unbounded. To verify
this claim, For positive integers n, let fn denote the element of c00 defined by
(
1 if m ≤ n
fn (m) =
0 if m > n.
Now ∥ιY,X (fn )∥1 = n, but ∥fn ∥∞ = 1.
(c) Consider c00 with the ∥ · ∥∞ norm. Let a : N → F be any function and define a linear
transformation Ta : c00 → c00 by
(5) Ta f (n) = a(n)f (n).
The mapping Ta is bounded if and only if M = supn∈N |a(n)| < ∞, in which case
∥Ta ∥ = M . In this case, Ta extends uniquely to a bounded operator from c0 to c0
17
by Proposition 1.33, and one may check that the formula (5) defines the extension.
All of these claims remain true if we use the ∥ · ∥1 norm instead of the ∥ · ∥∞ norm.
In this case, we get a bounded operator from ℓ1 to itself.
(d) Define S : ℓ1 → ℓ1 as follows: given the sequence (f (n))n from ℓ1 let Sf (1) = 0 and
Sf (n) = f (n − 1) for n > 1. (Viewing f as a sequence, S shifts the sequence one
place to the right and fills in a 0 in the first position). This S is an isometry, but is
not surjective. In contrast, if X is finite-dimensional, then the rank-nullity theorem
from linear algebra guarantees that every injective linear map T : X → X is also
surjective.
(e) Let C ∞ ([0, 1]) denote the vector space of functions on [0, 1] with continuous deriva-
tives of all orders. The differentiation map D : C ∞ ([0, 1]) → C ∞ ([0, 1]) defined by
df
Df = dx is a linear transformation. Since, for t ∈ R, we have Detx = tetx , it follows
that there is no norm on C ∞ ([0, 1]) such that dxd
is bounded.
1.5. Problems.
Problem 1.4. Prove equivalent norms define the same topology and the same Cauchy
sequences.
Problem 1.5. (a) Prove all norms on a finite dimensional vector space X are equivalent.
P P
Suggestion: Fix a basis e1 , . . . en for X and define ∥ ak ek ∥1 := |ak |. It is routine
to check that ∥ · ∥1 is a norm on X . Now complete the following outline.
(i) Let ∥ · ∥ be the given norm on X . Show there is an M such that ∥x∥ ≤ M ∥x∥1 .
Conclude that the mapping ι : (X , ∥ · ∥1 ) → (X , ∥ · ∥) defined by ι(x) = x is
continuous;
(ii) Show that the unit sphere S = {x ∈ X : ∥x∥1 = 1} in (X , ∥ · ∥1 ) is compact in
the ∥ · ∥1 topology;
(iii) Show that the mapping f : S → (X , ∥ · ∥) given by f (x) = ∥x∥ is continuous
and hence attains its infimum. Show this infimum is not 0 and finish the proof.
(b) Combine the result of part (a) with the result of Problem 1.4 to conclude that every
finite-dimensional normed vector space is complete.
(c) Let X be a normed vector space and M ⊆ X a finite-dimensional subspace. Prove
M is closed in X .
for all x, y ∈ [0, 1]. Define ∥f ∥Lip to be the best possible constant in this inequality.
That is,
|f (x) − f (y)|
∥f ∥Lip := sup
x̸=y |x − y|
Let Lip[0, 1] denote the set of all Lipschitz continuous functions on [0, 1]. Prove ∥f ∥ :=
|f (0)| + ∥f ∥Lip is a norm on Lip[0, 1], and that Lip[0, 1] is complete in this norm.
Problem 1.8. Let C 1 ([0, 1]) denote the space of all functions f : [0, 1] → R such that
f is differentiable in (0, 1) and f ′ extends continuously to [0, 1]. Prove
∥f ∥ := ∥f ∥∞ + ∥f ′ ∥∞
is a norm on C 1 ([0, 1]) and that C 1 is complete in this norm. Do the same for the norm
∥f ∥ := |f (0)| + ∥f ′ ∥∞ . (Is ∥f ′ ∥∞ a norm on C 1 ?)
Problem 1.9. Let (X, M ) be a measurable space. Let M (X) denote the (real) vector
space of all signed measures on (X, M ). Prove the total variation norm ∥µ∥ := |µ|(X)
is a norm on M (X), and M (X) is complete in this norm.
Problem 1.10. Prove, if X , Y are normed spaces, then the operator norm is a norm on
B(X , Y).
Problem 1.11. Prove c00 is dense in c0 and ℓ1 . (That is, given f ∈ c0 there is a sequence
fn in c00 such that ∥fn − f ∥∞ → 0, and the analogous statement for ℓ1 .) Using these
facts, or otherwise, prove that c00 is not dense in ℓ∞ . (In fact there exists f ∈ ℓ∞ with
∥f ∥∞ = 1 such that ∥f − g∥∞ ≥ 1 for all g ∈ c00 .)
Problem 1.12. Prove c00 is not complete in the ∥ · ∥1 or ∥ · ∥∞ norms. (After we have
studied the Baire Category theorem, you will be asked to prove that there is no norm
on c00 making it complete.)
Problem 1.13. Consider c0 and c00 equipped with the ∥ · ∥∞ norm. Prove there is no
bounded operator T : c0 → c00 such that T |c00 is the identity map. (Thus the conclusion
of Proposition 1.33 can fail if Y is not complete.)
Problem 1.14. Prove the ∥ · ∥1 and ∥ · ∥∞ norms on c00 are not equivalent. Conclude
from your proof that the identity map on c00 is bounded from the ∥ · ∥1 norm to the
∥ · ∥∞ norm, but not the other way around.
Problem 1.15. a) Prove f ∈ C0 (Rn ) if and only if f is continuous and lim|x|→∞ |f (x)| =
0. b) Let Cc (Rn ) denote the set of continuous, compactly supported functions on Rn .
Prove Cc (Rn ) is dense in C0 (Rn ) (where C0 (Rn ) is equipped with sup norm).
Problem 1.16. Prove, if X , Y are normed spaces and X is finite dimensional, then
every linear transformation T : X → Y is bounded. Suggestion: Let d denote the
dimension of X and let {e1 , . . . , ed } denote a basis. The function ∥ · ∥1 on X defined by
P P
∥ xj ej ∥1 = |xj | is a norm. Apply Problem 1.5.
19
Problem 1.20. Let X be a normed vector space and M a proper closed subspace.
Prove for every ϵ > 0, there exists x ∈ X such that ∥x∥ = 1 and inf y∈M ∥x − y∥ > 1 − ϵ.
(Hint: take any u ∈ X \ M and let a = inf y∈M ∥u − y∥. Choose δ > 0 small enough so
a u−v
that a+δ > 1 − ϵ, and then choose v ∈ M so that ∥u − v∥ < a + δ. Finally let x = ∥u−v∥ .)
Note that the distance to a (closed) subspace need not be attained. Here is an
example. Consider the Banach space C([0, 1]) (with the sup norm of course and either
real or complex valued functions) and the closed subspace
Z 1
T = {f ∈ C([0, 1]) : f (0) = 0 = f dt}.
0
Using machinery in the next section it will be evident that T is a closed subspace of
C([0, 1]). For now, it can be easily verified directly. Let g denote the function g(t) = t.
Verify that, for f ∈ T , that
Z Z
1
= g dt = (g − f ) dt ≤ ∥g − f ∥∞ .
2
In particular, the distance from g to T is at least 21 .
Note that the function h = x − 12 , while not in T , satisfies ∥g − h∥∞ = 12 .
On the other hand, for any ϵ > 0 there is an f ∈ T so that ∥g − f ∥∞ ≤ 12 + ϵ
(simply modify h appropriately). Thus, the distance from g to T is 12 . Now
R verify, using
the inequality above, that h is the only element of C([0, 1]) such that h dt = 0 and
∥g − h∥∞ = 12 .
Problem 1.21. Prove, if X is an infinite-dimensional normed space, then the unit ball
ball(X ) := {x ∈ X : ∥x∥ ≤ 1} is not compact in the norm topology. (Hint: use the
result of Problem 1.20 to construct inductively a sequence of vectors xn ∈ X such that
∥xn ∥ = 1 for all n and ∥xn − xm ∥ ≥ 12 for all m < n.)
Problem 1.22. (The quotient norm) Let X be a normed space and M a proper closed
subspace.
a) Prove the quotient norm is a norm.
b) Show that the quotient map x → x + M has norm 1. (Use Problem 1.20.)
c) Prove, if X is complete, so is X /M.
20
2.1. Examples. This subsection contains some examples of bounded linear functionals
and dual spaces.
Example 2.3. For each of the sequence spaces c0 , ℓ1 , ℓ∞ , for each n the map f → f (n) is
a bounded linear functional. That is, λn : X → F defined by λn (f ) = f (n) for f : N → F
in X , where X is any one of c0 , ℓ1 , ℓ∞ , is continuous since in each case it is immediate
that
|λn (f )| = |f (n)| ≤ ∥f ∥X .
Example 2.4. Given g ∈ ℓ1 , if f ∈ c0 , then
X∞ X∞
(6) |f (n)g(n)| ≤ ∥f ∥∞ |g(n)| = ∥g∥1 ∥f ∥∞ .
n=0 n=0
P∞
Thus n=0 f (n)g(n) converges and we obtain a functional Lg : c0 → F defined by
∞
X
(7) Lg (f ) := f (n)g(n).
n=0
The inequality of equation (6) says Lg is bounded (continuous) and ∥Lg ∥ ≤ ∥g∥1 . More-
over, it is immediate that Φ : ℓ1 → c∗0 defined by Φ(g) = Lg is bounded and linear and
∥Φ∥ ≤ 1. In fact, Φ is onto so that every bounded linear functional on c0 is of the form
Lg for some g ∈ ℓ1 .
21
Proof. We have already seen that each g ∈ ℓ1 gives rise to a bounded linear functional
Lg ∈ c∗0 via
∞
X
Lg (f ) := g(n)f (n),
n=0
that ∥Lg ∥ ≤ ∥g∥1 and the the mapping Φ is bounded and linear. We will prove simul-
taneously that this map is onto and that ∥Lg ∥ ≥ ∥g∥1 .
Let L ∈ c∗0 . We will first show that there is unique g ∈ ℓ1 so that L = Lg . Let
en ∈ c0 be the indicator function of n, that is
en (m) = δnm .
Define a function g : N → F by
g(n) = L(en ).
We claim that g ∈ ℓ1 and L = Lg . To see this, fix an integer N and define h = hN :
N → F by
(
g(n)/|g(n)| if n ≤ N and g(n) ̸= 0
h(n) =
0 otherwise.
Thus h = N
P
n=0 h(n)en . Further, by h ∈ c00 ⊆ c0 and ∥h∥∞ ≤ 1. Now
N
X N
X
|g(n)| = h(n)g(n) = L(h) = |L(h)| ≤ ∥L∥∥h∥ ≤ ∥L∥.
n=0 n=0
Proof. The proof follows the same lines as the proof of the previous proposition; the
details are left as an exercise. □
Remark 2.8. The same mapping g → Lg also shows that every g ∈ ℓ1 gives a bounded
linear functional on ℓ∞ , but it turns out these do not exhaust (ℓ∞ )∗ (see Problem 2.12).
22
2.2. Continuous linear functionals. For linear functionals, we can add to the list of
equivalent conditions of Proposition 1.31. In particular, the proof of the equivalence of
items (a) and (c) in Proposition 2.12 requires a map into the scalar field.
23
Proof. Items (a) and (b) are equivalent by Proposition 1.31 and it is evident that item (a)
implies item (c).
Suppose item (b) does not hold. Thus, there exists a sequence (fn ) from X such
that ∥fn ∥ ≤ 1, but |λ(fn )| ≥ n. Choose e ∈ X with λ(e) = 1 and let
fn
hk = e −
λ(fn )
and note that (hk ) converges to e but λ(hk ) = 0 for all k. Thus (hk ) is a sequence from
ker λ that converges to a point not in ker λ. Thus item (c) implies item (b).
Since ker λ ̸= V (since λ is not the zero map), item (c) implies item (d). Now
suppose item (c) does not hold. Thus there exists an f ̸∈ ker λ and a sequence (fn )
from ker λ that converges to f. Without loss of generality, λ(f ) = 1. Given g ∈ X , the
sequence
gn = (g − λ(g)f ) + λ(g) fn
converges to g and λ(gn ) = 0. Thus g ∈ ker λ and we conclude X = ker λ. □
Remark 2.13. Note that Proposition 2.12 remains true with λ−1 ({a}) in place of ker λ,
for any choice of a ∈ F. For instance, in the proof that item (c) implies item (b), simply
require λ(e) = a + 1 instead of λ(e) = a/ In the proof that item (d) implies item (c),
suppose the sequence (fn ) converges to f and λ(fn ) = a, but λ(fn ) = b ̸= a. In this
case, given g ∈ X , let gn = (g − cf ) + cfn < where c = a−λ(g)
a−b
. The details are left as an
exercise.
As a corollary, Proposition 2.12 extends to linear maps from a normed space X into
a finite dimensional normed space as an easy argument shows.
If X is infinite dimensional, the result is false. Just choose a basis B for X and
let B0 = {b1 , b2 , . . . } denote a countable subset of B. Define L : X → X by declaring
L(bn ) = n bn and L(b) = b for b ∈ B \ B0 and extend by linearity. Thus L is one-one so
that ker L = {0} is closed, but L is not bounded (and so not continuous). □
We close this subsection with the following result that should be compared with
item (a) from Subsection 1.4 followed by an example.
Proposition 2.14. If V is an infinite dimensional normed vector space, then there exists
a linear map f : V → F that is not continuous.
24
For a Banach space X , there are notions of a basis that reference the norm. For
instance, a Schauder basis for X is a sequence (en )∞ n=1 such that
P∞for each x ∈ X there
exists a unique choice of scalars xn ∈ F such that the series n=1 xn en converges to
x. Forgetting the norm structure, a Hamel basis B ⊆ X for X is a basis in the sense
of linear algebra. Explicitly, letting F00 (B) denote the functions a : B → F such that
ab = a(b) is zero for all but finitely many b ∈ B, the set B is a Hamel basis for X if for
each v ∈ X there exist is a unique function a ∈ F00 (B) such that
X finite
X
v= ab b = ab b.
b∈B b∈B
In this case any choice of c : B → F determines uniquely a linear functional λ : X → F
via the rule X
λ(v) = c b ab ,
b∈B
where cb = c(b). Often this process is described informally as: let λ(b) = c(b) and extend
by linearity. Finally, an argument using Zorn’s Lemma, which we will soon encounter
in the proof of the Hahn-Banach Theorem, shows that every vector space has a basis.
While it is true that every basis for a vector space V has the same cardinality, all that
we need to make sense of the statement V is an infinite dimensional vector space is the
fact that V has a basis that is infinite, then all bases for V are infinite, which is an
immediate consequence of the fact that all bases for a finite dimensional vector space
have the same cardinality. Thus, we can take the statement X is infinite dimensional
to mean that X has a Hamel basis B that contains a countable set B0 .
Proof of Proposition 2.14. Let B denote a Hamel basis for V. By assumption, B has a
countable subset B0 . Write B0 = {b1 , b2 , . . . } (so choose a bijection ψ : N → B0 ) and
assume, without loss of generality that ∥bj ∥ = 1. Let λ : V → F denote the linear
functional determined by λ(bj ) = j for bj ∈ B0 and λ(b) = 0 for b ∈ B \ B0 and observe
that λ is not bounded. □
Example 2.15. Let X denote an infinite dimensional normed vector space and suppose
f : X → F denote a discontinuous linear functional. An exercise shows that f −1 ({1}) =
X in addition to ker f = X (see Proposition 2.12). Let U = {f ≤ 0} ⊆ ker λ and note
V = X \ U = {f > 0} ⊇ f −1 ({1}). Now U and V are disjoint convex sets such that
X = U ∪ V and U = X = V .
2.3. The Hahn-Banach Extension Theorem. To state and prove the Hahn-Banach
Extension Theorem, we first work in the setting F = R, then extend the results to the
complex case.
Definition 2.16. Let X be a real vector space. A Minkowski functional is a function
p : X → R such that p(x + y) ≤ p(x) + p(y) and p(λx) = λp(x) for all x, y ∈ X and
nonnegative λ ∈ R.
25
The proof will invoke Zorn’s Lemma, a result that is equivalent to the axiom of
choice (as well as the well-ordering principle and the Hausdorff maximality principle).
A partial order ⪯ on a set S is a relation that is reflexive, symmetric and transitive;
that is, for all x, y, z ∈ S
(i) x ⪯ x,
(ii) if x ⪯ y and y ⪯ x, then x = y, and
(iii) if x ⪯ y and y ⪯ z, then x ⪯ z.
We call S, or more precisely (S, ⪯), a partially ordered set or poset. A subset T of
S is totally ordered , if for each x, y ∈ T either x ⪯ y or y ⪯ x. A totally ordered subset
T is often called a chain. An upper bound z for a chain T is an element z ∈ S such that
t ⪯ z for all t ∈ T . A maximal element for S is a w ∈ S that has no successor; that is
there does not exist an s ∈ S such that s ̸= w and w ⪯ s. An upper bound for a subset
A of S is an element s ∈ S such that a ⪯ s for all a ∈ A.
Theorem 2.19 (Zorn’s Lemma). Suppose S is a partially ordered set. If every chain in
S has an upper bound, then S has a maximal element.
In particular, for m ∈ M,
L(m)−λ ≤ p(m − x)
(8)
L(m)+λ ≤ p(m + x).
Let N = M + Rx and define L′ : N → R by L′ (m + tx) = L(m) + tλ for m ∈ M
and t ∈ R. Thus L′ is linear and agrees with L on M by definition. Moreover, by
construction and equation 8,
L(m − x) ≤ p(m − x)
(9)
L(m + x) ≤ p(m + x).
We now check that L′ (y) ≤ p(y) for all y ∈ M + Rx. Accordingly, suppose y ∈ N
so that there exists m ∈ M and t ∈ R such that y = m + tx. If t = 0 there is nothing
to prove. If t > 0, then, in view of the second inequality of equation (9),
m m
′ ′
L (y) = L (m + tx) = t L( ) + λ ≤ t p( + x) = p(m + tx) = p(y)
t t
and a similar estimate, using the first inequality of equation (9), shows that
L′ (m + tx) ≤ p(m + tx)
for t < 0. We have thus successfully extended L to a linear map L′ : N → R satisfying
L′ (n) ≤ p(n) for all n ∈ N and the proof is complete. □
We make one further observation before turning to the proof of the Hahn-Banach
Theorem. If T is a totally ordered set and (Nα )α∈T are subspaces of a vector space X
that are nested increasing in the sense that Nα ⊆ Nβ for α ⪯ β, then N = ∪α∈T Nα is
again a subspace of X . By contrast, if X is a normed vector space and Nα are (closed)
subspaces of X , then N will not necessarily be a (closed) subspace of X .
Proof of Theorem 2.17. Let L denote the set of pairs (L′ , N ) where N is a subspace of
X containing M, and L′ is an extension of L to N obeying L′ (y) ≤ p(y) on N . Declare
(L′1 , N1 ) ⪯ (L′2 , N2 ) if N1 ⊆ N2 and L′2 |N1 = L′1 . This relation ⪯ is a partial order on L;
that is (L, ⪯) is a partially ordered set. Further, Lemma 2.20 says if (L′ , N ) is maximal
element, then N = X .
An exercise shows, given any increasing chain (L′α , Nα ) in L has as an upper bound
(L′ , N ) in L, where N := α Nα and L′ (nα ) := L′α (nα ) for nα ∈ Nα . By Zorn’s Lemma
S
27
the collection L has a maximal element (L′ , N ) with respect to the order ⪯ and the
proof is complete. □
The use of Zorn’s Lemma in the proof of Theorem 2.17 is a typical - one knows how
to carry out a construction one step a time, but there is no clear way to do it all at once.
As an exercise, use Zorn’s Lemma to prove that if V Is a vectors space and S ⊆ V is a
linearly independent set, then there is a basis B for V such that B ⊇ S.
In the special case that p is a seminorm, since L(−x) = −L(x) and p(−x) = p(x)
the inequality L ≤ p is equivalent to |L| ≤ p.
Corollary 2.21. Suppose X is a real normed vector space, M is a subspace of X , and
L is a bounded linear functional on M. If C ≥ 0 and |L(x)| ≤ C∥x∥ for all x ∈ M,
then there exists a bounded linear functional L′ on X extending L such that ∥L′ ∥ ≤ C.
Proof. Apply the Hahn-Banach theorem with the Minkowski functional p(x) = C∥x∥.
□
(i) L′ |M = L and
(ii) |L′ (x)| ≤ p(x) for all x ∈ X .
Sketch of proof. Using the Lemma 2.22 and its notation: The proof consists of applying
the real Hahn-Banach theorem (Corollary 2.21) to the R-linear functional u = ReL to
obtain a real linear functional u′ : X → R extending u and satisfying u′ (x) ≤ p(x) for all
x ∈ X . The resulting complex functional L′ associated to u′ is then a desired extension
of L. The details are left as an exercise. □
The following corollaries are quite important, and when the Hahn-Banach theorem
is applied it is usually in one of the following forms:
Corollary 2.25. Let X be a normed vector space over F (either R or C).
(i) If M ⊆ X is a subspace and L : M → F is a bounded linear functional, then there
exists a bounded linear functional L′ : X → F such that L′ |M = L and ∥L′ ∥ = ∥L∥.
(ii) (Linear functionals detect norms) If x ∈ X is nonzero, there exists L ∈ X ∗ with
∥L∥ = 1 such that L(x) = ∥x∥.
(iii) (Linear functionals separate points) If x ̸= y in X , there exists L ∈ X ∗ such that
L(x) ̸= L(y).
(iv) (Linear functionals detect distance to subspaces) If M ⊆ X is a closed subspace
and x ∈ X \ M, there exists L ∈ X ∗ such that
(a) L|M = 0;
(b) ∥L∥ = 1; and
(c) L(x) = dist(x, M) = inf y∈M ∥x − y∥ > 0.
(v) if L is a linear submanifold of X and x ∈ X , then x ∈ L if and only if λ(x) = 0
for every λ ∈ X ∗ for which L ⊆ ker λ.
Proof. To prove item (i) consider the (semi)norm p(x) = ∥L∥ ∥x∥. By construction,
|L(x)| ≤ p(x) for x ∈ M. Hence, there is a linear functional L′ on X such that
L′ |M = L and |L′ (x)| ≤ p(x) for all x ∈ X . In particular, ∥L′ ∥ ≤ ∥L∥. On the other
hand, ∥L′ ∥ ≥ ∥L∥ since L′ agrees with L on M.
For item (ii), let M be the one-dimensional subspace of X spanned by x. Define a
x
functional L : M → F by L(t ∥x∥ ) = t. In particular, |L(y)| = ∥y∥ for y ∈ M and thus
∥L∥ = 1. By (i), the functional L extends to a functional (still denoted L) on X such
that ∥L∥ = 1.
An application of item (ii) to the vector x − y proves item (iii).
To prove item (iv), let δ = dist(x, M). Since M is closed, δ > 0. Define a functional
L : M + Fx → F by L(y + tx) = tδ for y ∈ M and t ∈ F. Since for t ̸= 0 and y ∈ M,
∥y + tx∥ = |t|∥t−1 y + x∥ ≥ |t|δ = |L(y + tx)|,
by Hahn-Banach we can extend L to a functional L ∈ X ∗ with ∥L∥ ≤ 1.
29
2.4. The bidual and reflexive spaces. Note that since X ∗ is a normed space, we
can form its dual, denoted X ∗∗ , and called the bidual or double dual of X . There is
a canonical relationship between X and X ∗∗ . Each fixed x ∈ X gives rise to a linear
functional x̂ : X ∗ → F via evaluation,
x̂(L) := L(x).
Since |x̂(L)| = |L(x)| ≤ ∥L∥ ∥x∥, the linear functional x̂ is in X ∗∗ and ∥x̂∥ ≤ ∥x∥.
Corollary 2.26. (Embedding in the bidual) The map x → x̂ is an isometric linear map
from X into X ∗∗ .
The embedding into the bidual has many applications; one of the most basic is the
following.
Proposition 2.31 (Completion of normed spaces). If X is a normed vector space, then
there is a Banach space X and an isometric linear map ι : X → X such that the image
ι(X ) is dense in X .
Proof. Embed X into X ∗∗ via the map x → x̂ and let X be the closure of the image of
X in X ∗∗ . Since X is a closed subspace of a complete space, it is complete. □
31
2.5. Dual spaces and adjoint operators. [Optional] Let X , Y be normed spaces
with duals X ∗ , Y ∗ . If T : X → Y is a linear transformation and f : Y → F is a linear
functional, then T ∗ f : X → F defined by
(10) (T ∗ f )(x) = f (T x)
is a linear functional on X . If T and f are both continuous (that is, bounded) then the
composition T ∗ f is bounded, and more is true:
2.6. Duality for Sub and Quotient Spaces. [Optional. Not covered Spring 2025]
The Hahn-Banach Theorem allows for the identification of the duals of subspaces and
quotients of Banach spaces. Informally, the dual of a subspace is a quotient and the dual
of a quotient is a subspace. The precise results are stated below for complex scalars, but
they hold also for real scalars.
32
Given a (closed) subspace M of the Banach space X , let π denote the map from X
to the quotient X /M. Recall (see Problem 1.22), the quotient is a Banach space with
the norm,
∥z∥ = inf{∥y∥ : π(y) = z}.
In particular, if x ∈ X , then
∥π(x)∥ = inf{∥x − m∥ : m ∈ M}.
It is evident from the construction that π is continuous and ∥π∥ ≤ 1. Further, by
Problem 1.20 (or see Proposition 2.33 below) if M is a proper (closed) subspace, then
∥π∥ = 1. In particular, π ∗ : (X /M)∗ → X ∗ (defined by π ∗ λ = λ ◦ π) is also continuous.
Moreover, if x ∈ M, then
π ∗ λ(x) = λ(π(x)) = 0.
Let
M⊥ = {f ∈ X ∗ : f (x) = 0 for all x ∈ M}.
(M⊥ is called the annihilator of M in X ∗ .) Recall, given x ∈ X , the element x̂ ∈ X ∗∗
is defined by x̂(τ ) = τ (x), for τ ∈ X ∗ . In particular,
M⊥ = ∩x∈M ker(x̂)
and thus M⊥ is a closed subspace of X ∗ . Further, if λ ∈ (X /M)∗ , then π ∗ λ ∈ M⊥ .
Proposition 2.33 (The dual of a quotient). The mapping ψ : (X /M)∗ → M⊥ defined
by
ψ(λ) = π ∗ λ
is an isometric isomorphism; i.e., the mapping π ∗ : (X /M)∗ → X ∗ is an isometric
isomorphism onto M⊥ .
Proof. The linearity of ψ follows from Theorem 2.32 as does ∥ψ∥ = ∥π∥ ≤ 1. To prove
that ψ is isometric, let λ ∈ (X /M)∗ be given. Automatically, ∥ψ(λ)∥ ≤ ∥λ∥. To prove
the reverse inequality, fix r > 1. Let q ∈ X /M with ∥q∥ = 1 be given. There exists an
x ∈ X such that ∥x∥ < r and π(x) = q. Hence,
|λ(q)| = |λ(π(x))∥ = ∥ψ(λ)(x)∥ ≤ ∥ψ(λ)∥ ∥x∥ < r∥ψ(λ)∥.
Taking the supremum over such q shows ∥λ∥ ≤ r∥ψ(λ)∥. Finally, since 1 < r is arbitrary,
∥λ∥ ≤ ∥ψ(λ)∥.
To prove that ψ is onto, and complete the proof, let τ ∈ M⊥ be given. Fix q ∈ X /M.
If x, y ∈ X and π(x) = q = π(y), then τ (x) = τ (y). Hence, the mapping λ : X /M → C
defined by λ(q) = τ (x) is well defined. That λ is linear is left as an exercise. To see that
λ is continuous, observe that
|λ(q)| = |τ (x)| ≤ ∥τ ∥ ∥x∥,
33
for each x ∈ X such that π(x) = q. Taking the infimum over such x gives shows
|λ(q)| ≤ ∥τ ∥ ∥q∥.
Finally, by construction ψ(λ) = τ. □
Proof. It remains to show that φ is an isometry, a fact that is an easy consequence of the
Hahn-Banach Theorem. Fix λ ∈ M∗ and let q = φ(λ). If f is any bounded extension
of λ to X ∗ , then ∥f ∥ ≥ ∥λ∥. Hence,
∥φ(λ)∥ =∥q∥
= inf{∥f ∥ : f ∈ X ∗ , ρ(f ) = q}
= inf{∥f ∥ : f ∈ X ∗ , f |M = λ}
≥∥λ∥.
On the other hand, by the Hahn-Banach Theorem there is a bounded extension g of λ
with ∥g∥ = ∥λ∥. Thus ∥λ∥ ≤ ∥q∥. □
A special case of the following useful fact was used in the proofs above. If X , Y are
vector spaces and T : X → Y is linear and M is a subspace of the kernel of T , then T
induces a linear map Te : X /M → Y. A canonical choice is M = ker(T ) in which case
Te is one-one. If X is a Banach space, Y is a normed vector space and M is closed, then
X /M is a Banach space.
Proof. Let π : X → X /M denote the quotient map and observe that Teπ = T . Since the
quotient map π has norm 1 (see Problem 1.22), we see that ∥Te∥ ≤ ∥T ∥. For the opposite
34
inequality, let 0 < ϵ < 1 and choose x ∈ X such that ∥x∥ = 1 and ∥T x∥ > (1 − ϵ)∥T ∥.
Then ∥π(x)∥ ≤ 1 and
∥T̃ ∥ ≥ ∥T̃ π(x)∥ = ∥T x∥ > (1 − ϵ)∥T ∥.
Letting ϵ go to zero finishes the proof. □
Proof. It is easy to check that the closure of subspace of X is again a subspace. It follows
that M is a subspace with M ⊆ M ⊆ X . Since M has codimension 1, we must have
either M = M or M = X . □
Proposition 2.38. Let X be a normed vector space and L : X → F a linear functional.
Then L is continuous (that is, bounded) if and only if ker L is closed. Consequenlty, L
is continuous if and only if there exists a nonempty open set U such that U ∩ ker L = ∅.
Recall that a set K in a real vector space X is called convex if for every x, y ∈ K and
every 0 ≤ t ≤ 1, we have tx + (1 − t)y ∈ K. Let X be a normed vector space over R and
let U ⊆ X be a convex, open set containing 0. We define a function p : X → [0, +∞) by
1
(12) p(x) = inf{r > 0 : x ∈ U }.
r
(To see that the definition makes sense, observe that since U is open and 0 ∈ U , there
exists δ > 0 so that x ∈ U whenever ∥x∥ < δ. It follows that for every x ∈ X , we
have 1r x ∈ U for all r > ∥x∥
δ
; thus the set appearing in the definition is nonempty.) The
functional p is sometimes called a gauge for the set U , it is important because of the
following lemma.
Lemma 2.39. Let X be a normed vector space over R and let U ⊆ X be a convex open
set containing 0. Then the function p defined by (12) is a Minkowski functional, and
U = {x ∈ X : p(x) < 1}.
Proof. We first assume that 0 ∈ U , the general case will follow by translation. Let
N be the one-dimensional subspace Rx. Define L on N by putting L(x) = 1 and
extending linearly. Let p be the gauge functional for U . Since x ∈
/ U , we have p(x) ≥ 1,
so 1 = L(x) ≤ p(x). Since both L and p are positive homogeneous, we also have
L(tx) ≤ p(tx) for all t ≥ 0. For t < 0, we have L(tx) < 0 ≤ p(tx) (since p ≥ 0 by
definition). Thus, we have L(y) ≤ p(y) for all y in the subspace N . It follows from the
Hahn-Banach theorem that there exists an extension of L to all of X (still denoted L)
36
such that L(y) ≤ p(y) for all y ∈ X . It follows that L(x) = 1 and L(y) ≤ p(y) < 1 for
all y ∈ U . To see that this extension L is bounded, let V = U − x; then V is an open
set in X and L(y) < 0 for all y ∈ V , that is, V ∩ ker L = ∅, so by the second part of
Proposition 2.38 it follows that L is bounded.
Finally, in the case that U does not contain 0, we choose a point x0 ∈ U and apply
the theorem to U ′ := U − x0 and x′ = x − x0 , to obtain a bounded functional such that
L(x) = 1 + L(x0 ) and L(y) < 1 + L(x0 ) for all y ∈ U ; the details are left to the reader.
□
Lemma 2.41. i) If X is a normed vector space over R, K ⊆ X is a convex set,
and x0 is an interior point of K, then for every x ∈ K and every 0 ≤ t < 1, the
point x0 + t(x − x0 ) is an interior point of K.
ii) If K is convex then int(K) is convex.
iii) If K is a closed, convex subset of X and K has nonempty interior, then K is
equal to the closure of its interior.
iv) Let K be a closed, convex subset of X and suppose 0 is an interior point of K. If
p is the gauge functional for the convex set U = int(K), then K = {x ∈ X |p(x) ≤
1}.
Proof. Again we assume that 0 is an interior point of K, and leave the general case to the
reader. Let U = int(K) and let p be the gauge functional for U ; by the lemma we have
K = {x ∈ X |p(x) ≤ 1}. Thus, if x ∈/ K, then p(x) > 1. We may then choose a number
0 < t < 1 so that tx ∈
/ K. Applying the previous separation theorem to U and tx, we
obtain a bounded linear functional L and a real number a such that L(y) < a = L(tx)
37
for all y ∈ U , we put b := at = L(x). Since L(y) < a on U , and L is continuous, and
K is the closure of U (by item (iii) of the Lemma), we conclude that L(y) ≤ a for all
a ∈ K, which completes the proof. □
2.8. Problems.
Problem 2.1. Prove, if X is any normed vector space, {x1 , . . . xn } is a linearly indepen-
dent set in X , and α1 , . . . αn are scalars, then there exists a bounded linear functional f
on X such that f (xj ) = αj for j = 1, . . . n. (Recall linear maps from a finite dimensional
normed vector space to a normed vector space are bounded.)
Problem 2.2. Let X , Y be normed spaces and T : X → Y a linear transformation.
Prove T is bounded if and only if there exists a constant C such that for all x ∈ X and
f ∈ Y ∗,
(13) |f (T x)| ≤ C∥f ∥∥x∥;
in which case ∥T ∥ is equal to the best possible C in (13).
Problem 2.3. Let X be a normed vector space. Show that if M is a closed subspace of
X and x ∈ / M, then M + Fx is closed. Use this result to give another proof that every
finite-dimensional subspace of X is closed.
Problem 2.4. Prove, if M is a finite-dimensional subspace of a Banach space X , then
there exists a closed subspace N ⊆ X such that M ∩ N = {0} and M + N = X . (In
other words, every x ∈ X can be written uniquely as x = y + z with y ∈ M, z ∈ N .)
Hint: Choose a basis x1 , . . . xn for M and construct, using Problem 2.1 and the Hahn-
Banach Theorem, bounded linear functionals f1 , . . . fn on X such that fi (xj ) = δij . Now
let N = ∩ni=1 ker fi . (Warning: this conclusion can fail badly if M is not assumed finite
dimensional, even if M is still assumed closed. Perhaps the first known example is that
c0 is not complemented in ℓ∞ , though it is nontrivial to prove.)
Problem 2.5. Prove Proposition 2.22.
Problem 2.6. Prove every finite-dimensional Banach space is reflexive.
Problem 2.7. Let B denote the subset of ℓ∞ consisting of sequences which take values
in {−1, 1}. Show that any two (distinct) points of B are a distance 2 apart. Show, if C
is a countable subset of ℓ∞ , then there exists a b ∈ B such that ∥b − c∥ ≥ 1 for all c ∈ C.
Conclude ℓ∞ is not separable. Prove there is no isometric isomorphism Λ : c0 → ℓ∞ .
Problem 2.8. [This problem belongs in the section with signed measures] Prove, if µ
is a finite regular (signed) Borel measure on a compact Hausdorff space, then the linear
function Lµ : C(X) → R defined by
Z
Lµ (f ) = f dµ
X
38
is bounded (continuous) and ∥Lµ ∥ = ∥µ∥ := |µ|(X). (See the Riesz-Markov Theorem
for positive linear functionals.)
Problem 2.9. Let X and Y be normed vector spaces and T ∈ L(X , Y).
a) Consider T ∗∗ : X ∗∗ → Y ∗∗ . Identifying X , Y with their images in X ∗∗ and Y ∗∗ ,
show that T ∗∗ |X = T .
b) Prove T ∗ is injective if and only if the range of T is dense in Y.
c) Prove that if the range of T ∗ is dense in X ∗ , then T is injective; if X is reflexive
then the converse is true.
d) Assuming now that X and Y are Banach spaces, prove that T : X → Y is
invertible if and only if T ∗ is invertible, in which case (T ∗ )−1 = (T −1 )∗ .
Problem 2.12. a) Prove there exists a bounded linear functional L ∈ (ℓ∞ )∗ with
the following property: whenever f ∈ ℓ∞ and limn→∞ f (n) exists, then L(f ) is
equal to this limit. (Hint: first show that the set of such f forms a subspace
M ⊆ ℓ∞ ).
b) Show that such a functional L is not equal to Lg for any g ∈ ℓ1 ; thus the map
T : ℓ1 → (ℓ∞ )∗ given by T (g) = Lg is not surjective.
c) Give another proof that T is not surjective, using Problem 2.11.
Problem 2.13. Let X be a normed space and let K ⊆ X be a convex set. (Recall,
this means that whenever x, y ∈ K, then 12 (x + y) ∈ K; equivalently, tx + (1 − t)y ∈ K
for all 0 ≤ t ≤ 1.) A point x ∈ K is called an extreme point of K whenever y, z ∈ K,
0 < t < 1, and x = ty + (1 − t)z, then y = z = x. (That is, the only way to write x as a
convex combination of elements of K is the trivial way.)
a) Let X be a normed space and let B = ball(X ) denote the (closed) unit ball of
X . Prove that x ∈ B is not an extreme point of B if and only if there exists a
nonzero y ∈ B such that ∥x ± y∥ ≤ 1.
39
This section contains three important applications of the Baire category theorem in
functional analysis. These are the Principle of Uniform boundedness (also known as the
Banach-Steinhaus theorem), the Open Mapping Theorem, and the Closed Graph The-
orem. (In learning these theorems, keep careful track of what completeness hypotheses
are needed.)
Theorem 3.2 (The Baire Category Theorem). Suppose X is a complete metric space.
(a) If (Un )∞ ∞
n=1 is a sequence of open dense subsets of X, then ∩n=1 Un ̸= ∅.
(b) If (Fn )n is a sequence of closed sets with empty interior, then ∪Fn ̸= X.
Remark 3.3. We will actually prove that ∩Un is dense in X. This conclusion is in fact
equivalent to the conclusion that ∩Un ̸= ∅.
Theorem 3.2 is true if X is a locally compact Hausdorff space and there are connec-
tions between the Baire Category Theorem and the axiom of choice. □
The following lemma should be familiar from advanced calculus. It will be used in
the proof of Theorem 3.2.
Lemma 3.4. Let X be a complete metric space and suppose (Cn ) is a sequence of subsets
of X. If
(i) each Cn is nonempty;
(ii) (Cn ) is nested decreasing;
(iii) each Cn is closed; and
(iv) (diam(Cn )) converges to 0,
then there is an x ∈ X such that
{x} = ∩Cn .
Moreover, if (xn ) is a sequence from X and xn ∈ Cn for each n, then (xn ) converges to
some x.
Proof of Theorem 3.2. The two items are equivalent, but for our purposes it is enough to
show item (a) implies item (b). To this end, suppose item (a) holds, (Fn ) is a sequence
of closed sets such that Fn◦ = ∅ for all n and X = ∪Fn . Taking complements, ∅ = ∩Fnc .
By Lemma 3.1, the sets Fnc are open and dense in X by Lemma 3.1. Hence X is not
complete and therefore item (a) implies item (b). Thus it suffices to prove item (a).
To prove item (a), let (Un )∞
n=1 be a sequence of open dense sets in X and let I = ∩Un .
To prove I is dense, it suffices to show that I has nontrivial intersection with every
nonempty open set W by Lemma 3.1. Fix such a W . Since, by Lemma 3.1, U1 is dense,
41
there is a point x1 ∈ W ∩ U1 . Since U1 and W are open, there is a radius 0 < r1 < 1 such
that the B(x1 , r1 ) is contained in W ∩ U1 by Lemma 3.1. Similarly, since U2 is dense
and open there is a point x2 ∈ B(x1 , r1 ) ∩ U2 and a radius 0 < r2 < 12 such that
B(x2 , r2 ) ⊆ B(x1 , r1 ) ∩ U2 ⊆ W ∩ U1 ∩ U2 .
Continuing inductively, since each Un is dense and open there is a sequence of points
(xn )∞ 1
n=1 and radii 0 < rn < n such that
The sequence of sets (B(xn , rn )) satisfies the hypothesis of Lemma 3.4 and X is complete.
Hence there is an x ∈ X such that
x ∈ ∩n B(xn , rn ) ⊆ W ∩ I. □
3.2. Category. Baire’s theorem is used as a kind of pigeonhole principle: the “thick”
complete metric space X cannot be expressed as a countable union of “thin” closed sets
without interior.
Definition 3.5. A subset E of a metric space X is nowhere dense (in X) if its closure
has empty interior; that is (E)◦ = ∅.
A set F in a metric space X is first category (or meager ) if it can be expressed as a
countable union of nowhere dense sets. In particular, a countable union of first category
sets is first category.
A set G is second category if it is not first category.
Corollary 3.6 (The Baire Category Theorem restated). If X is a complete metric space,
then X is not a countable union of nowhere dense sets; that is, X is of second category
in itself.
Corollary 3.7. An infinite dimensional Banach space can not have a countable basis.
In particular, there is no norm on c00 that makes it a Banach space; ditto for the vector
space of polynomials.
Proof. The proof is left as an exercise. As a starting point, show, if M is finite dimen-
sional subspace of an infinite dimensional Banach space X , then M is nowhere dense in
X. □
42
If B is of the second category (thus not a countable union of nowhere dense sets) in X,
then
sup ∥Tα ∥ < ∞.
α
Corollary 3.9. Suppose X is a Banach space and Y is a normed vector space. If (Tn )
is a sequence of bounded operators Tn : X → Y that converges pointwise to a (necessarily
linear) map T : X → Y, then T is bounded.
Outline of proof. In a metric space, convergent sequences are bounded. Hence (Tn x)n is
bounded in Y for each x ∈ X . Thus the set X = {x ∈ X : sup{∥Tn x∥ : n ∈ N} < ∞}
is of second category in X . Thus C = sup{∥Tn ∥ : n ∈ N} < ∞. Thus the proof reduces
to showing ∥T x∥ ≤ C∥x∥ for all x ∈ X , a task that is left as an exercise for the gentle
reader. □
43
3.4. Open mapping. Given a subset B of a vector space X and a scalar s ∈ F, let
sB = {sb : b ∈ B}. Similarly, for x ∈ X , let B − x = {b − x : b ∈ B}. Let X , Y be
normed vector spaces and suppose T : X → Y is linear. If B ⊆ X and s ∈ F is nonzero,
then T (sB) = sT (B) and further, an easy argument shows T (sB) = s T (B). It is also
immediate that if B is open, then so is B − x.
Recall that if X, Y are topological spaces, a mapping f : X → Y is called open if
f (U ) is open in Y whenever U is open in X. In particular, if f is a bijection, then f is
open if and only if f −1 is continuous. In the case of normed linear spaces the condition
that a linear map is open is refined by the Open Mapping Theorem, Theorem 3.11 below.
But first a lemma.
Lemma 3.10 (Translation and Dilation lemma). Let X , Y be normed vector spaces, let
B denote the open unit ball of X , and let T : X → Y be a linear map. The following are
equivalent.
(i) The map T is open;
(ii) T (B) contains an open ball centered at 0;
(iii) there is an s > 0 such that T (sB) contains an open ball centered at 0; and
(iv) T (sB) contains an open ball centered at 0 for each s > 0.
In the proof of this lemma and that of Theorem 3.11 to follow, we use B X (x, r) and
B Y (y, s) to denote the open balls centered to x and y with radii r and s in X and Y
respectively when needed to avoid ambiguity.
Proof. This result is more or less immediate from the fact that, for fixed z0 ∈ X and
r ∈ F, the translation map z → z + z0 and the dilation map z → rz are continuous in a
normed vector space.
The implication item (i) implies item (ii) is immediate. The fact that T (sB) =
sT (B) for s > 0 readily shows items (ii), (iii) and (iv) are all equivalent.
To finish the proof it suffice to show item (iv) implies item (i). Accordingly, suppose
item (iv) holds and let U ⊆ X be a given open set. To prove that T (U ) is open, let
y ∈ T (U ) be given. There is an x ∈ U such that T (x) = y. There is an s > 0 such that
the ball B(x, s) lies in U ; that is B(x, s) ⊆ U . The ball B(0, s) = B(x, s) − x is an open
ball centered to 0. By hypothesis there is an r > 0 such that B Y (0, r) ⊆ T (B(0, s)). By
linearity of T ,
B Y (y, r) =B Y (0, r) + y ⊆ T (B(0, s)) + y
=T (B(0, s)) + T (x) = T (B(0, s) + x) = T (B(x, s)) ⊆ T (U ).
Thus T (U ) is open and the proof is complete. □
(i) T (X ) = Y;
(ii) Y is complete (so a Banach space); and
(iii) T is open.
In particular, if X , Y are Banach spaces, and T : X → Y is bounded and onto, then
T is an open map.
Proof. Assuming T is open and letting B denote the open unit ball in X , by Lemma 3.10,
there is an r > 0 such that B Y (0, r) ⊆ T (B). Hence,
Y = ∪∞ Y
n=1 B (0, n r) ⊆ T (nB) ⊆ T (X ),
so that item (i) holds. (Here the superscript Y is used to emphasize this ball is in Y.)
That is, item (iii) implies item (i).
To prove item S∞ r centered at x in X .
S∞ (iii), let B(x, r) denote the open ball of radius
Trivially X = n=1 B(0, n) and thus range T = T (X ) = n=1 T (B(0, n)). Since the
range of T is assumed second category, there is an N such that T (B(0, N )) is second
category and hence somewhere dense. In other words, T (B(0, N )) has nonempty interior.
By scaling (see Lemma (3.10)), T (B(0, 1)) has nonempty interior. Hence, there exists
p ∈ Y and r > 0 such that T (B(0, 1)) contains the open ball B Y (p, r). It follows that
for all ∥y∥ < r,
y = (y + p) − p ∈ T (B(0, 1)) + T (B(0, 1)) ⊆ T (B(0, 2)),
where we have used −T (B(0, 1)) = T (B(0, 1)). In other words,
B Y (0, r) ⊆ T (B(0, 2)).
By scaling, it follows that, for n ∈ N,
r 1
B Y (0, ) ⊆ T (B(0, )).
2n+1 2n
END Friday 2025-01-31
We will use the hypothesis that X is complete to prove B Y (0, 4r ) ⊆ T (B(0, 1)), which,
by Lemma 3.10, implies T is open. Accordingly let y such that ∥y∥ < 4r be given. Since
y is in the closure of T (B(0, 12 )), there is a y1 ∈ T (B(0, 21 )) such that ∥y − y1 ∥ < 8r . Since
y − y1 ∈ B Y (0, 8r ) it is is in the closure of T (B(0, 41 )). Thus there is a y2 ∈ T (B(0, 41 ))
such that ∥(y − y1 ) − y2 ∥ < 16 r
. Continuing in this fashion produces a sequence (yj )∞ j=1
from Y such that,
(a) ∥y − nj=1 yj ∥ ≤ 2n+2 r
P
; and
1
(b) yn ∈ T (B(0, 2n ))
for all n. It follows the sequence (sm = m
P
j=1 yj )m converges to y. P Further, for each
j there is an xj ∈ B(0, 2j ) such that yj = T xj . Thus, setting tm = m
1
j=1 xj , we have
45
T sm = tm and
∞
X ∞
X
∥xj ∥ < 2−k = 1,
j=1 k=1
P∞
thus the sequence (tm )m converges to some x with ∥x∥ ≤ j=1 ∥xj ∥ < 1, that is,
x ∈ B(0, 1). It follows that y = T x by continuity of T . Consequently y ∈ T (B(0, 1))
and the proof of item (iii) is complete.
To prove item (ii), let M denote the kernel of T and Te the mapping Te : X /M → Y
determined by Teπ = T ; that is Te(π(x)) = T (x) for x ∈ X . By construction Te is one-one
and by Lemma 2.35, it is continuous. Further its range is the same as the range of T ,
namely Y, and is thus second category. Hence, by what has already been proved, Te is
an open map. and consequently Te−1 is continuous. Hence X /M and Y are isomorphic
(though of course not necessarily isometrically isomorphic) as normed vector spaces.
Therefore, since X /M is complete (Proposition 1.28), so is Y. □
Note that the proof of item (ii) in the Open Mapping Theorem shows, in the case
that in the case that T is one-one and its range is of second category, that T is onto and
its inverse is continuous. In particular, if T : X → Y is a continuous bijection and Y is
a Banach space (so the range of T is second category), then T −1 is continuous.
Corollary 3.12 (The Banach Isomorphism Theorem). If X , Y are Banach spaces and
T : X → Y is a bounded bijection, then T −1 is also bounded (hence, T is an isomor-
phism).
Proof sketch. Note that when T is bijection, T is open if and only if T −1 is continuous.
The result thus follows from the Open Mapping Theorem and Proposition 1.31. □
The following examples show that the hypothesis that X and Y are Banach spaces,
and not just normed vector spaces, is needed in Corollary 3.12.
Example 3.13. This example shows that the assumption that the range of T is second
category in Y is necessary in Theorem 3.11.
Let X denote the Banach space ℓ1 and let Y = ℓ1 as a linear manifold in c0 with
the c0 (sup) norm. So Y is a normed space, but not a Banach space. The identity map
ιX toY is a bijection. It is also continuous since the supremum norm of an element of ℓ1
dominates its ℓ1 norm. Let G = B X (0, 1) ⊆ X denote the (open) unit ball in X. Thus
G is open in X . Given r > 0 choose n ∈ N such that n > 2r and let x = 2r nj=1 ej , where
P
Observe that ∥x∥∞ < r, but ∥x∥1 > 1. Hence x ∈ B Y (0, r), but x ∈/ G. Thus B Y (0, r) ̸⊆
G for any choice of r > 0, which means 0 is not an interior point of G (in Y). Hence G
is not open in Y.
The argument above of course shows that the range of ι is not second category Pn in
1
Y. Here is simple direct proof of this fact. First note that the sequence gn = n j=1 ej
converges to 0 in c0 and gn ∈ ℓ1 with ∥gn ∥1 = 1. For N ∈ N, let BN = {f ∈ ℓ1 : ∥f ∥1 ≤
N } ⊆ Y. Verify that each BN is closed in Y. On the other hand, given N, the sequence
fk = f + 3N gk converges to f in c0 and so in Y, but ∥f + 3N gk ∥ ≥ 3N − ∥f ∥ ≥ 2N > N.
Thus f is not in the interior of BN and so BN is nowhere dense and Y = ∪∞ N =1 BN .
Example 3.14. This example shows that the assumption X is a Banach space can not
be relaxed to X is simply a normed vector space in Theorem 3.11.
Let Y be an infinite dimensional Banach space. Let λ be a discontinuous linear
functional, whose existence is the content of Proposition 2.14. As an exercise, show that
the function ∥ · ∥∗ : Y → [0, ∞) given by
∥x∥∗ = ∥x∥Y + |λ(x)|
is a norm on Y. Let X denote the normed space Y with this norm; that is X = (Y, ∥·∥∗ ).
Let T : X → Y denote the identity map (so bijective). Let G denote the unit ball in X .
We claim 0 is not in the interior of G as a subset of Y. Indeed, since λ is not continuous,
there is a sequence (xn ) of unit vectors in Y such that |λ(x)| ≥ n. Consequently, given
r > 0 and choosing n sufficiently large, ∥ 2r xn ∥Y < r, but ∥ 2r xn ∥∗ > 1. Thus B(0, r) ̸⊆ G
proving the claim.
A consequence of the argument is that X is not a Banach space. To verify this fact
directly, let y ∈
/ ker λ be given. By Proposition prop:bdd-lf-iff-cns, there is a sequence
(xn ) from ker λ that converges to x (in Y). Since ∥xn ∥∗ = ∥x∥Y , the sequence (xn ) is
Cauchy in X . However, since ∥xn − y∥∗ = ∥xn − y∥Y − λ(y), the sequence (xn ) does
not converge to y in X . Now suppose (xn ) converged to some z ∈ X . Thus z ∈ Y and
∥xn − z∥∗ = ∥xn − z∥Y + λ(z) converges to 0 from which it follows that z = y and the
proof is complete.
This result depends on the axiom of choice. In this proof, choice is smuggle in
through the appeal to Propostion 2.14, whose proof in turn depended on the existence
of a Hamel basis, which in turn uses Zorn’s Lemma (choice). □
END Monday 2025-02-03 - except had not discussed the Banach Isomorphism The-
orem.
3.5. The Closed Graph Theorem. Recall that the Cartesian product X × Y of Ba-
nach spaces X and Y with its default product topology from Subsection 1.2.7. In
particular, the product topology on X × Y is the coarsest topology that makes both
coordinate projections πX and πY from X × Y to X and Y respectively continuous. This
topology is the same as that determined by the norms in equation (2).
47
Definition 3.15. The graph of a linear map T : X → Y between normed vector spaces
is the set
G(T ) := {(x, y) ∈ X × Y : y = T x}.
Observe that since T is a linear map, G(T ) is a linear subspace of X × Y. The transfor-
mation T is closed if G(T ) is a closed subset of X × Y. □
It is an easy exercise to show that G(T ) is closed if and only if whenever (xn , T xn )
converges to (x, y), we have y = T x. Problem 3.2 gives an example where G(T ) is closed,
but T is not continuous. On the other hand, the next theorem says that if X , Y are
complete (Banach spaces), then G(T ) is closed if and only if T is continuous.
Theorem 3.16 (The Closed Graph Theorem). If X , Y are Banach spaces and T : X →
Y is closed, then T is bounded.
Proof. We prove T closed implies T is bounded, leaving the easy converse as an exercise.
With the norm ∥(x, y)∥∞ = max{∥x∥, ∥y∥} the vector space X ×Y is a Banach space with
the product topology. The coordinate projections πX , πY are bounded with norm one.
Let π1 , π2 be the coordinate projections πX , πY restricted to G(T ); explicitly π1 (x, T x) =
x and π2 (x, T x) = T x. Note that π1 is a bijection between G(T ) and X and in particular
π1−1 (x) = (x, T x). By hypothesis G(T ) is a closed subset of a Banach space and hence
a Banach space. Thus π1 is a bounded linear bijection between Banach spaces and
therefore, by Corollary 3.12, π1−1 : X → G(T ) is bounded. Since π2 is bounded, π2 ◦π1−1 :
X → Y is continuous. To finish the proof, observe π2 ◦ π1−1 (x) = π2 (x, T x) = T x. □
3.6. Problems.
Problem 3.1. Show that there exists a sequence of open, dense subsets Un ⊆ R such
that m( ∞
T
n=1 Un ) = 0.
Problem 3.5. Let X , Y be Banach spaces. Suppose (Tn ) is a sequence in B(X , Y) and
limn Tn x exists for every x ∈ X . Prove, if T is defined by T x = limn Tn x, then T is
bounded.
Problem 3.6. Suppose that X is a vector space with a countably infinite basis. (That
is, there is a linearly independent set {xn } ⊆ X such that every vector x ∈ X is
expressed uniquely as a finite linear combination of the xn ’s.) Prove there is no norm
on X under which it is complete. (Hint: consider the finite-dimensional subspaces
Xn := span{x1 , . . . xn }.)
Problem 3.7. The Baire Category Theorem can be used to prove the existence of
(very many!) continuous, nowhere differentiable functions on [0, 1]. To see this, let En
denote the set of all functions f ∈ C[0, 1] for which there exists x0 ∈ [0, 1] (which may
depend on f ) such that |f (x) − f (x0 )| ≤ n|x − x0 | for all x ∈ [0, 1]. Prove the sets En
are nowhere dense in C[0, 1]; the Baire Category Theorem then shows that the set of
nowhere differentiable functions is second category. (To see that En is nowhere dense,
approximate an arbitrary continuous function f uniformly by piecewise linear functions
g, whose pieces have slopes greater than 2n in absolute value. Any function sufficiently
close to such a g will not lie in En .)
Problem 3.8. Let L2 ([0, 1]) denote the Lebesgue measurable functions f : [0, 1] → C
such that |f |2 is in L1 ([0, 1]). It turns out, as we will see later, that L2 ([0, 1]) is a linear
manifold (subspace of the vector space L1 ([0, 1])), though this fact is not needed for this
problem.
Let gn : [0, 1] → R denote the function which takes the value n on [0, n13 ] and 0
elsewhere. Show,
R
(i) if f ∈ L2 ([0, 1]), then limn→∞ gn f dm R= 0;
(ii) Ln : L1 ([0, 1]) → C defined by Ln (f ) = gn f dm is bounded, and ∥Lg ∥ = n;
(iii) conclude L2 ([0, 1]) is of the first category in L1 ([0, 1]).
Problem 3.10. Suppose X is a Banach space and M and N are closed subspaces.
Show, if for each x ∈ X there exist unique m ∈ M and n ∈ N such that
x = m + n,
then the mapping P : X → M defined by P x = m is bounded.
49
4. Lp spaces
Proposition 4.4. The set L∞ (µ) is a vector space. Further, the infimum in equa-
tion (15) is attained and ∥ · ∥∞ is a semi-norm on L∞ (µ).
We record the following simple observation for later use - often without comment.
4.1. Conjugate indices and the inequalities of Young, Holder and Minkowski.
We now restrict our attention to 1 ≤ p ≤ ∞.
Definition 4.6. The conjugate index or dual exponent to 1 < p < ∞ is the unique
1 < q < ∞ satisfying
1 1
+ = 1.
p q
The dual index to p = ∞ is q = 1; and the dual index to p = 1 is q = ∞. □
Note that (p − 1)q = p and likewise (q − 1)p = q. The significance of dual indices is
apparent in the following result.
Lemma 4.7 (Young’s inequality). If a, b are nonnegative numbers and 1 < p, q < ∞
are dual indices, then
ap bq
ab ≤ +
p q
q p
and equality holds if and only if b = a .
If µ has the property that every set of positive measure contains a set of positive,
but finite, measure and f ∈ L∞ (µ), then
Remark 4.9. For f ∈ L1 (µ) of course equality holds in equation (17) with g = 1.
The assumption that (X, M , µ) has the property that every (measurable) set of
positive measure contains a set of finite measure is needed as the following example
shows. For the measure space ({0}, {∅, {0}}, µ), where µ(∅) = 0 and µ({0}) = ∞, we
have L1 (µ) = {0} and thus ∥f g∥1 = 0 for all f ∈ L∞ (µ) and g ∈ L1 (µ). □
Proof of Theorem 4.8. The proof is easy in the cases p = ∞ or p = 1. Now suppose
1 < p < ∞.
If ∥f ∥p = 0, then f = 0 a.e. by Lemma 4.5. Hence f g = 0 a.e. and, by another
application of Lemma 4.5, ∥f g∥1 = 0. Thus the inequality of equation (17) holds. By
symmetry, the same is true for g. Hence we may assume ∥f ∥p ̸= 0 ̸= ∥g∥q .
By homogeneity we may assume ∥f ∥p = ∥g∥q = 1. We are to show
Z
|f g| dµ ≤ 1.
Example 4.10. One can get a more intuitive feel for what Hölder’s inequality says by
examining it in the case of step functions. Let E, F be sets of finite, positive measure
and put f = 1E , g = 1F . Then ∥f g∥1 = µ(E ∩ F ) and
∥f ∥p ∥q∥q = µ(E)1/p µ(F )1/q ,
1
so Hölder’s inequality can be proved easily in this case using the relation p
+ 1q = 1 and
the fact that µ(E ∩ F ) ≤ min(µ(E), µ(F )).
Corollary 4.11 (Minkowski’s inequality). Let (X, M , µ) be a measure space and sup-
pose 1 ≤ p ≤ ∞. If f, g ∈ Lp (µ), then
∥f + g∥p ≤ ∥f ∥p + ∥g∥p .
Hence ∥ · ∥p is a semi-norm on Lp (µ).
Proof. The result has already been established for p = 1 and p = ∞ so suppose 1 <
p < ∞ and let q denote the conjugate index to p. By Lemma 4.2, f + g ∈ Lp (µ). The
result is vacuous if f + g = 0 (almost everywhere); equivalenlty, ∥f + g∥p = 0. Now
suppose ∥f + g∥p ̸= 0. By Theorem 4.8, there is an h ∈ Lq (µ) such that ∥h∥q = 1 and
∥(f + g)h∥1 = ∥f + g∥p . On the other hand, two more applications of Theorem 4.8 and
the fact that ∥ · ∥1 is a semi-norm gives,
∥f + g∥p = ∥(f + g)h∥1 ≤ ∥f h∥1 + ∥gh∥1 ≤ ∥f ∥p ∥h∥1 + ∥g∥p ∥h∥1 = ∥f ∥p + ∥g∥p . □
4.2. The Lebesuge spaces Lp (µ). The proof of the following proposition, based on
Lemma 4.5 is left to the gentle reader.
Proposition 4.12. The set N (ν) = {f ∈ Lp (µ) : ∥f ∥p = 0} is a subspace of Lp (µ) and
the function ∥ · ∥p descends to a norm on the quotient space Lp (µ)/N (µ).
Definition 4.13. The normed vector space (Lp (µ)/N (µ), ∥ · ∥p ) (for 1 ≤ p ≤ ∞) is
denoted Lp (µ) and is known as a Lebesgue space.
then Φ is isometric. The proof of this claim is left as an (easy) exercise for the gentle
reader.
Later we will see that the map Φ in Proposition 4.14 is an isometric isomorphism
for 1 ≤ p < ∞, with the proviso that µ is σ-finite in the case p = 1.
Proof of Proposition 4.14. As already observed, ∥Ψ(g)∥ = ∥Lg ∥ ≤ ∥g∥q , for g ∈ Lq (µ).
On the other hand, since 1 ≤ q < ∞, for 0 ̸= g ∈ Lq (µ), the moreover portion of Hölder’s
Inequality (Theorem 4.8) gives a function f ∈ Lp (µ) such that ∥f ∥p = 1 and
Z
Lg (f ) = f g = ∥g∥q .
Proposition 4.16. For 1 ≤ p ≤ ∞, the normed vector space Lp (µ) is a Banach space.
Sketch of proof. The proof for the case 1 ≤ p < ∞ is very much like the case p = 1 that
has already been established and is just sketched here.
There is a subsequence (gk )∞
k=1 of (fn ) such that ∥gk+1 −gk ∥ < 2
−k
for k ≥ 1. Setting
P∞
g0 = 0, the series k=0 ∥gk+1 − gk ∥p converges. (The subsequence (gk ) is super-Cauchy.)
Let
Xm
hm = |gk+1 − gk |
k=0
and let h denote the pointwise limit (in [0, ∞]) of the non-negative increasing sequence
(hm ). By the Monotonne Convergence Theorem,
Z Z
(20) h = lim hpm .
p
The inequality,
m
" ∞
#p
X X
∥hm ∥p ≤ ∥gk+1 − gk ∥p ≤ ∥gk+1 − gk ∥p <∞
k=0 k=0
55
also converges almost everywhere to a measurable function f. That is, the sequence (gk )
converges pointwise to f. Further, since |f | ≤ h and h ∈ Lp , it follows that f ∈ Lp .
By construction, for m fixed, if n > m, then ∥gn − gm ∥p < 21−m and (|gn − gm |p )n
converges pointwise almost everywhere to |f − gm |p . Thus, by Fatou,
Z Z
∥f − gm ∥p = |f − gm | ≤ lim inf |gn − gm |p = lim inf ∥gn − gm ∥pp < 21−m .
p p
Proof. The sequence (fn ) (viewed as representative of their respective equivalence classes)
satisfies the hypotheses of Lemma 4.17. Hence there is a g ∈ Lp (µ) and subsequence (gk )
of (fn ) that converges to g both in Lp (µ) and pointwise almost everywere. By uniqueness
of limits, g = f as elements of Lp (µ); that is, almost everywhere. Thus (gk ) converge to
f pointwise almost everywhere. □
Use the Closed Graph Theorem as follows. Suppose (fn , Mh fn ) is a sequence that
converges to (f, g) in Lp ×Lr . Apply Corollary 4.18 to (fn ) and f to obtain a subsequence
(gk ) of (fn ) converging to f pointwise a.e. and of course in Lp . Apply Corollary 4.18 to
(hgk ) and g to deduce g = hf. Now use Closed Graph.
For the bit about L2 , make an argument like the one at the end of the proof of
Hölder’s inequality.
4.3. Problems.
Problem 4.1. Suppose f : [0, A] → [0, ∞) is differentiable, strictly increasing and
f (0) = 0. Prove, for each 0 < a ≤ A, that
Z x Z f (x)
f + f −1 = xf (x).
0 0
Rx R f (x)
[Suggestion: Differentiate g(x) = 0
f + 0
f −1 − xf (x).] Deduce Young’s inequality.
Problem 4.2. [Truncation of Lp functions] Suppose f is an unsigned function in Lp (µ),
1 < p < ∞. For t > 0 let
Et = {x : f (x) > t}.
Show:
(a) For each real number t > 0, the horizontal truncation 1Et f belongs to Lq for all
1 ≤ q ≤ p.
(b) For each real number t > 0, the vertical truncation ft := min(f, t) belongs to Lq for
all p ≤ q ≤ ∞.
(c) Every f ∈ Lp , 1 < p < ∞, can be decomposed as f = g + h where g ∈ L1 and
h ∈ L∞ .
Problem 4.3. Suppose f ∈ Lp0 ∩L∞ for some p0 < ∞. Prove f ∈ Lp for all p0 ≤ p ≤ ∞,
and limp→∞ ∥f ∥p = ∥f ∥∞ .
Problem 4.4. Prove fn → f in the L∞ norm if and only if fn → f essentially uniformly,
and that L∞ is complete.
Problem 4.5. Show that Lp (R) ̸⊆ Lq (R) for any pair p, q.
Problem 4.6. Consider L∞ (R).
a) Show that M := C0 (R) is a closed subspace of L∞ (R) (more precisely, that the
set of L∞ functions that are a.e. equal to a C0 function is closed in L∞ ). Prove
there is a bounded linear functional λ : L(R)∞ → F such that λ|M = 0 and
λ(1R ) = 1.
b) Prove there is no function g ∈ L1 (R) such that λ(f ) = R f g dm for all f ∈ L∞ .
R
weakly bounded, 47
Zorn’s Lemma, 25