6617notes24jan14
6617notes24jan14
SPRING 2014
19.1. Examples.
1/2
a) Of course, Rn with the usual Euclicean norm k(x1 , . . . xn )k = ( nk=1 |xk |2 )
P
is a
n n
Banach space. (Likewise C with the Euclidean norm.) Besides these, K can be
equipped with the `p -norms
n
!1/p
X
k(x1 , . . . xn )kp := |xk |p (19.4)
k=1
It is not too hard to show that all of the `p norms (1 ≤ p ≤ ∞) are equivalent on Kn
(though the constants c, C must depend on the dimension n). 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 20.3.
b) (Sequence spaces) Define
c0 := {f : N → K| lim |f (n)| = 0}
n→∞
∞
` := {f : N → K| sup |f (n)| < ∞}
n∈N
X∞
`1 := {f : N → K| |f (n)| < ∞}
n=0
2
It is a simple exercise to check that each of these is a vector space. Define for
functions f : N → K
kf k∞ := sup |f (n)|
n
∞
X
kf k1 := |f (n)|.
n=1
XN Z XN
f− fn = f− fn dm (19.8)
n=1 X n=1
1
Z ∞
X
≤ |fn | dm (19.9)
X n=N +1
X
= kfn k1 → 0 (19.10)
n=N +1
3
as N → ∞. (What justifies the equality in the last line?)
d) (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
kf kp := |f | dm <∞ (19.11)
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). The sequence
space `p is defined analagously: it is the set of f : N → K for which
∞
!1/p
X
kf kp := |f (n)|p <∞ (19.12)
n=1
as for the other Lp spaces we identify f and g when there are equal a.e. When
f ∈ L∞ , let kf k∞ be the smallest M for which (19.13) holds. Then k · k∞ is a norm
making L∞ (m) into a Banach space.
e) (C(X) spaces) Let X be a compact metric space and let C(X) denote the set of
continuous functions f : X → K. It is a standard fact from advanced calculus that
the quantity kf k∞ := supx∈X |f (x)| is a norm on C(X). A sequence is Cauchy in
this norm if and only if it is uniformly Cauchy. It is thus also a standard fact that
C(X) is complete in this norm—completeness just means that a uniformly Cauchy
sequence of continuous functions on X converges uniformly to a continuous function.
This example can be generalized somewhat: let X be a locally compact metric
space. Say a function f : X → K vanishes at infinity if for every > 0, there
exists a compact set K ⊂ X such that supx∈K / |f (x)| < . Let C0 (X) denote the
set of continuous functions f : X → K that vanish at inifinity. Then C0 (X) is a
vector space, the quantity kf k∞ := supx∈X |f (x)| is a norm on C0 (X), and C0 (X) is
complete in this norm. (Note that c0 from above is a special case.)
f) (Subspaces and direct sums) If (X , k · k) is a normed vector space and Y ⊂ X is a
vector subspace, then the restriction of k · k to Y is clearly a norm on Y. If X is a
Banach space, then (Y, k · k) is a Banach space if and only if Y is closed in the norm
topology of X . (This is just a standard fact about metric spaces—a subspace of a
complete metric space is complete in the restricted metric if and only if it is closed.)
If X , Y are vector spaces then the algebraic direct sum is the vector space of ordered
pairs
X ⊕ Y := {(x, y) : x ∈ X , y ∈ Y} (19.14)
4
with entrywise operations. If X , Y are equipped with norms k · kX , k · kY , then each
of the quanitites
k(x, y)k∞ := max(kxkX , kykY ),
k(x, y)k1 := kxkX + kykY
is a norm on X ⊕ Y. These two norms are equivalent; indeed it follows from the
definitions that
k(x, y)k∞ ≤ k(x, y)k1 ≤ 2k(x, y)k∞ . (19.15)
If X and Y are both complete, then X ⊕ Y is complete in both of these norms. The
resulting Banach spaces are denoted X ⊕∞ Y, X ⊕1 Y respectively.
g) (Quotient 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. If M is a closed subspace of X , then the quantity
kx + Mk := inf kx − yk (19.16)
y∈M
is a norm on X /M, called the quotient norm. (Geometrically, kx+Mk is the distance
in X from x to the closed set M.) It turns out that if X is complete, so is X /M.
See Problem 20.20.
More examples are given in the exercises. Shortly we will construct further examples from
linear transformations T : X → Y; to do this we first need to build up a few facts.
two useful notions of equivalence for normed vector spaces. Two normed spaces X , Y are said
to be (boundedly) isomorphic if there exists an invertible linear transformation T : X → Y.
The spaces are isometrically isomorphic if additionally kT xk = kxk for all x. A T with
this property is called an isometry; note that an isometry automatically injective; if it is
also surjective then it is automatically invertible, in which case T −1 is also an isometry. An
isometry need not be surjective, however.
19.3. 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.
b) Let X be c00 equipped with the k · k1 norm, and Y be c00 equipped with the k · k∞
norm. Then the identity map id : c00 → c00 is bounded as an operator from X to Y
(in fact its norm is equal to 1), but is unbounded as an operator from Y to X .
c) Consider c00 with the k · k∞ norm. Let a : N → K be any function and define a linear
transformation Ta : c00 → c00 by
Ta f (n) = a(n)f (n). (19.26)
Then Ta is bounded if and only if M = supn∈N |a(n)| < ∞, in which case kTa k = M .
When this happens, Ta extends uniquely to a bounded operator from c0 to c0 , and
one may check that the formula (19.26) defines the extension. All of these claims
7
remain true if we use the k·k1 norm instead of the k·k∞ norm; we then get a bounded
operator from `1 to itself.
d) For f ∈ `1 , define the shift operator by 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 space of functions on [0, 1] with continuous derivatives of all
df
orders. The differentiation map f → dx is a linear transformation from C ∞ [0, 1] to
e = tetx for all real t, we find that there is no norm on C ∞ [0, 1] for
d tx
itself. Since dx
d
which dx is bounded.
20. Problems
Problem 20.1. Prove Proposition 19.2.
Problem 20.2. Prove that equivalent norms define the same topology and the same Cauchy
sequences.
Problem 20.3. a) Prove that all norms on a finite dimensional P space X are
vector P
equivalent. (Hint: fix a basis e1 , . . . en for X and define k ak ek k1 := |ak |. Com-
pare any given norm k · k to this one. Begin by proving that the “unit sphere”
S = {x ∈ X : kxk1 = 1} is compact in the k · k1 topology.)
b) Combine the result of part (a) with the result of Problem 20.2 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
that M is closed in X .
Problem 20.4. Finish the proofs from Example 19.1(b).
Problem 20.5. A function f : [0, 1] → K is called Lipschitz continuous if there exists a
constant C such that
|f (x) − f (y)| ≤ C|x − y| (20.1)
for all x, y ∈ [0, 1]. Define kf kLip to be the best possible constant in this inequality. That is,
|f (x) − f (y)|
kf kLip := sup (20.2)
x6=y |x − y|
Let Lip[0, 1] denote the set of all Lipschitz continuous functions on [0, 1]. Prove that kf k :=
|f (0)| + kf kLip is a norm on Lip[0, 1], and that Lip[0, 1] is complete in this norm.
Problem 20.6. 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 0 extends continuously to [0, 1]. Prove that
kf k := kf k∞ + kf 0 k∞ (20.3)
is a norm on C 1 [0, 1] and that C 1 is complete in this norm. Do the same for the norm
kf k := |f (0)| + kf 0 k∞ . (Is kf 0 k∞ a norm on C 1 ?)
Problem 20.7. Let (X, M ) be a measurable space. Let M (X) denote the (real) vector
space of all signed measures on (X, M ). Prove that the total variation norm kµk := |µ|(X)
is a norm on M (X), and M (X) is complete in this norm.
8
Problem 20.8. Prove that if X , Y are normed spaces, then the operator norm is a norm on
B(X , Y).
Problem 20.9. Prove that c00 is dense in c0 and `1 . (That is, given f ∈ c0 there is a
sequence fn in c00 such that kfn − f k∞ → 0, and the analogous statement for `1 .) Using
these facts, or otherwise, prove that c00 is not dense in `∞ . (In fact there exists f ∈ `∞ with
kf k∞ = 1 such that kf − gk∞ ≥ 1 for all g ∈ c00 .)
Problem 20.10. Prove that c00 is not complete in the k · k1 or k · k∞ 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 20.11. Consider c0 and c00 equipped with the k · k∞ norm. Prove that there is no
bounded operator T : c0 → c00 such that T |c00 is the identity map. (Thus the conclusion of
Proposotion 19.9 can fail if Y is not complete.)
Problem 20.12. Prove that the k · k1 and k · k∞ norms on c00 are not equivalent. Conclude
from your proof that the identity map on c00 is bounded from the k · k1 norm to the k · k∞
norm, but not the other way around.
Problem 20.13. a) Prove that 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
that Cc (Rn ) is dense in C0 (Rn ) (where C0 (Rn ) is equipped with sup norm).
Problem 20.14. Prove that if X , Y are normed spaces and X is finite dimensional, then
every linear transformation T : X → Y is bounded.
Problem 20.15. Prove the claims in Example 19.3(c).
Problem 20.16. Let g : R → K be a (Lebesgue) measurable function. The map M g :
f → gf is a linear transformation on the space of measurable functions. Prove that Mg is
bounded from L1 (R) to itself if and only if g ∈ L∞ (R), in which case kMg k = kgk∞ .
Problem 20.17. Prove the claims about direct sums in Example 19.1(f).
Problem 20.18. Let X be a normed vector space and M a proper closed subspace. Prove
that for every > 0, there exists x ∈ X such that kxk = 1 and inf y∈M kx − yk > 1 − .
(Hint: take any u ∈ X \ M and let a = inf y∈M ku − yk. Choose δ > 0 small enough so that
a u−v
a+δ
> 1 − , and then choose v ∈ M so that ku − vk < a + δ. Finally let x = ku−vk .)
Problem 20.19. Prove that if X is an infinite-dimensional normed space, then the unit ball
ball(X ) := {x ∈ X : kxk ≤ 1} is not compact in the norm topology. (Hint: use the result of
Problem 20.18 to construct inductively a sequence of vectors xn ∈ X such that kxn k = 1 for
all n and kxn − xm k ≥ 21 for all m < n.)
Problem 20.20. (The quotient norm) Let X be a normed space and M a proper closed
subspace.
a) Prove that the quotient norm is a norm (see Example 19.1(g)).
b) Show that the quotient map x → x + M has norm 1. (Use Problem 20.18.)
c) Prove that if X is complete, so is X /M.
9
Problem 20.21. A normed vector space X is called separable if it is separable as a metric
space (that is, there is a countable subset of X which is dense in the norm topology). Prove
that c0 and `1 are separable, but `∞ is not. (Hint: for `∞ , show that there is an uncountable
collection of elements {fα } such that kfα − fβ k = 1 for α 6= β.)
21. Linear functionals and the Hahn-Banach theorem
If there is a “fundamental theorem of functional analysis,” it is the Hahn-Banach theorem.
The particular version of it we will prove is somewhat abstract-looking at first, but its
importance will be clear after studying some of its corollaries.
Let X be a normed vector space over the field K. A linear functional on X is a linear map
L : X → K. As one might expect, we are especially interested in bounded linear functionals.
Since K = R or C is complete, the vector space of bounded linear functionals B(X , K) is
itself a normed vector space, and is always complete (even if X is not). This space is called
the dual space of X and is denoted X ∗ . It is not yet obvious that X ∗ need be non-trivial
(that is, that there are any bounded linear functionals on X besides 0). One corollary of
the Hahn-Banach theorem will be that there always exist many linear functionals on any
normed space X (in particular, enough to separate points of X ).
21.1. Examples.
a) For each of the sequence spaces c0 , `1 , `∞ , for each n the map f → f (n) is a bounded
linear functional. If we fix g ∈ `1 , then the functional Lg : c0 → K defined by
X ∞
Lg (f ) := f (n)g(n) (21.1)
n=0
is bounded, since
∞
X ∞
X
|Lg (f )| ≤ |f (n)g(n)| ≤ kf k∞ |g(n)| = kgk1 kf k∞ . (21.2)
n=0 n=0
This shows that kLg k ≤ kgk1 . In fact, equality holds, and every bounded linear
functional on c0 is of this form:
Proposition 21.1. The map g → Lg is an ismoetric isomorphism from `1 onto the
dual space c∗0 .
Proof. We have already seen that each g ∈ `1 gives rise to a bounded linear functional
Lg ∈ c∗0 via
∞
X
Lg (f ) := g(n)f (n) (21.3)
n=0
and that kLg k ≤ kgk1 . We will prove simultaneously that this map is onto and that
kLg k ≥ kgk1 .
Let L ∈ c∗0 , we will first show that there is unique g ∈ `1 so that L = Lg . Let
en ∈ c0 be the indicator function of n, that is
en (m) = δnm . (21.4)
Define a function g : N → K by
g(n) = L(en ). (21.5)
10
We claim that g ∈ `1 and L = Lg . To see this, fix an integer N and let h ∈ c00 be
the function
(
g(n)/|g(n)| if n ≤ N and g(n) 6= 0
h(n) = . (21.6)
0 otherwise
N
X N
X
|g(n)| = h(n)g(n) = L(h) = |L(h)| ≤ kLkkhk ≤ kLk. (21.7)
n=0 n=0
It follows that g ∈ `1 and kgk1 ≤ kLk. Moreover, the same calculation shows
that L = Lg when restricted to c00 , so by the uniqueness of extensions of bounded
operators, L = Lg . Uniqueness of g is clear from its construction, since if L = Lg , we
must have g(n) = Lg (en ). Thus the map g → Lg is onto and kLg kc0 ∗ = kgk1 .
Proposition 21.2. (`1 )∗ is isometrically isomorphic to `∞ .
Proof. The proof follows the same lines as the proof of the previous proposition; the
details are left as an exercise.
The same mapping g → Lg also shows that every g ∈ `1 gives a bounded linear
functional on `∞ , but it turns out these do not exhaust (`∞ )∗ (see Problem 22.7).
b) If f ∈ L1 (m) and g is a bounded measurable function with supx∈X |g(x)| = M , then
the map
Z
Lg (f ) := f g dm (21.8)
X
is a bounded linear functional of norm at most M . We will prove in Section ?? that
the norm is in fact equal to M , and every bounded linear functional on L1 (m) is of
this type (at least when m is σ-finite).
c) If X is a compact metric space and µ is a finite, signed Borel measure on X, then
Z
Lµ (f ) := f dµ (21.9)
X
and define L0 (m + tx) = L(m) + tλ. This L0 is linear by definition and agrees with L on
M. We now check that L0 (y) ≤ p(y) for all y ∈ M + Rx. This is immediate if y ∈ M. In
general let y = m + tx with m ∈ M and t > 0. Then
m m m m
L0 (m + tx) = t L( ) + λ ≤ t L( ) + p( + x) − L( ) = p(m + tx) (21.14)
t t t t
and a similar estimate shows that L0 (m + tx) ≤ p(m + tx) for t < 0.
We have thus successfully extended L to M + Rx. To finish, let L denote the set of pairs
(L0 , N ) where N is a subspace of X containing M, and L0 is an extension of L to N obeying
L0 (y) ≤ p(y) on N . Declare (L01 , N1 ) ≤ (L02 , N2 ) if N1 ⊂ N2 and L02 |nn1 = L01 . This is a
partial order on L. SGiven any increasing chain (L0α , Nα ) in L, it has an upper bound (L0 , N )
in L, where N := α Nα and L(nα ) := L0α (nα ) for nα ∈ Nα . By Zorn’s lemma, then, the
collection L has a maximal element (L0 , N ) with respect to the order ≤. Since it always
possible to extend to a strictly larger subspace, the maximal element must have N = X ,
and the proof is finished.
The proof is a typical application of Zorn’s lemma—one knows how to carry out a con-
struction one step a time, but there is no clear way to do it all at once.
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 21.5. Let X be a normed vector space over R, M a subspace, and L a bounded
linear functional on M satisfying |L(x)| ≤ Ckxk for all x ∈ M. Then there exists a bounded
linear functional L0 on X extending L, with kL0 k ≤ C.
Proof. Apply the Hahn-Banach theorem with the Minkowski functional p(x) = Ckxk.
Before obtaining further corollaries, we extend these results to the complex case. First,
if X is a vector space over C then trivially it is also a vector space over R, and there is a
simple relationship between the R- and C-linear functionals.
12
Proposition 21.6. Let X be a vector space over C. If L : X → C is a C-linear functional,
then u(x) = ReL(x) defines an R-linear functional on X , and L(x) = u(x) − iu(ix). Con-
versely if u : X → R is R-linear then L(x) := u(x) − iu(ix) is C-linear. If in addition X is
normed, then kuk = kLk.
Proof. Problem ??
Theorem 21.7 (The Hahn-Banach Theorem, complex version). Let X be a normed vector
space over C, p a seminorm on X , M a subspace of X , and L : M → C a C-linear functional
satisfying |L(x)| ≤ p(x) for all x ∈ M. Then there exists a linear functional L0 : X → C
satisfying |L0 (x)| ≤ p(x) for all x ∈ X .
Proof. The proof consists of applying the real Hahn-Banach theorem to extend the R-linear
functional u = ReL to a functional u0 : X → R and then defining L0 from u0 as in Proposi-
tion 21.6. 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 21.8. Let X be a normed vector space.
(i) (Linear functionals detect norms) If x ∈ X is nonzero, there exists L ∈ X ∗ with
kLk = 1 such that L(x) = kxk.
(ii) (Linear functionals separate points) If x 6= y in X , there exists L ∈ X ∗ such that
L(x) 6= L(y).
(iii) (Linear functionals detect distance to subspaces) If M ⊂ X is a closed subspace
and x ∈ X \ M, there exists L ∈ X ∗ such that L|M = 0 and L(x) = dist(x, M) =
inf y∈M kx − yk > 0.
Proof. (i): Let M be the one-dimensional subspace of X spanned by x. Define a functional
x
L : M → K by L(t kxk ) = t. For p(x) = kxk, we have |L(x)| ≤ p(x) on M, and L(x) = kxk.
By the Hahn-Banach theorem the functional L extends to a functional (still denoted L) on
X and satisfies |L(y)| ≤ kyk for all y ∈ X ; thus kLk ≤ 1; since L(x) = kxk by construction
we conclude kLk = 1.
(ii): Apply (i) to the vector x − y.
(iii): Let δ = dist(x, M). Define a functional L : M + Kx → K by L(y + tx) = tδ. Since
for t 6= 0
ky + txk = |t|kt−1 y + xk ≥ |t|δ = |L(y + tx)|, (21.15)
∗
by Hahn-Banach we can extend L to a functional L ∈ X with kLk ≤ 1.
Needless to say, the proof of the Hahn-Banach theorem is thoroughly non-constructive,
and in general it is an important (and often difficult) problem, given a normed space X , to
find some concrete description of the dual space X ∗ . Usually this means finding a Banach
space Y and a bounded (or, better, isometric) isomorphism T : Y → X ∗ .
A final corollary, which is also quite important. Note that since X ∗ is a normed space, we
can form its dual, denoted X ∗∗ and called the bidual of X . There is a canonical relationship
between X and X ∗∗ . First observe that if L ∈ X ∗ , each fixed x ∈ X gives rise to a linear
functional x̂ : X ∗ → K via evaluation:
x̂(L) := L(x). (21.16)
13
Corollary 21.9. (Embedding in the bidual) The map x → x̂ is an isometric linear map from
X into X ∗∗ .
Proof. First, from the definition we see that
|x̂(L)| = |L(x)| ≤ kLkkxk (21.17)
so x̂ ∈ X ∗∗ and kx̂k ≤ kxk. It is straightforward to check (recalling that the L’s are linear)
that the map x → x̂ is linear. Finally, to show that kx̂k = kxk, fix a nonzero x ∈ X . From
Corollary 21.8(i) there exists L ∈ X ∗ with kLk = 1 and L(x) = kxk. But then for this x and
L, we have |x̂(L)| = |L(x)| = kxk so kx̂k ≥ kxk, and the proof is complete.
Definition 21.10. A Banach space X is called reflexive if the mapˆ: X → X ∗∗ is surjective.
In other words, X is reflexive if the mapˆ is an isomorphism of X with X ∗∗ . For example,
every finite dimensional Banach space is reflexive (Problem ??). On the other hand, we will
see below that c∗∗ ∞ p p
0 = ` , so c0 is not reflexive. After we have studied the L and ` spaces
p
in more detail, we will see that L is reflexive for 1 < p < ∞.
The embedding into the bidual has many applications; one of the most basic is the follow-
ing:
Proposition 21.11 (Completion of normed spaces). If X is a normed vector space, there is
a Banach space X and in isometric 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.
The space X is called the completion of X . It is unique in the sense that if Y is another
Banach space and j : X → Y embeds X isometrically as a dense subspace of Y, then Y is
isometrically isomorphic to X . The proof of this fact is left as an exercise.
21.2. Dual spaces and adjoint operators. Let X, Y be normed spaces with duals X ∗ , Y ∗ .
If T : X → Y is a linear transformation, then given any linear map f : Y → K we can define
another linear map T ∗ f : X → K by the formula
(T ∗ f )(x) = f (T x). (21.18)
In fact, if f is bounded then so is T ∗ f , and more is true:
Theorem 21.12. The formula (21.18) defines a bounded linear transformation T ∗ : Y ∗ →
X ∗ , and in fact kT ∗ k = kT k.
Proof. Let f ∈ Y ∗ and x ∈ X. Then
|T ∗ f (x)| = |f (T x)| ≤ kf kkT xk ≤ kf kkT kkxk. (21.19)
Taking the supremum over kxk = 1, we find that T ∗ f is bounded and kT ∗ f k ≤ kT kkf k,
so kT ∗ k ≤ kT k. For the reverse inequality, let 0 < < 1 be given and choose x ∈ X with
kxk = 1 and kT xk > (1 − )kT k. Now consider T x. By the Hahn-Banach theorem, there
exists f ∈ Y ∗ such that kf k = 1 and f (T x) = kT xk. For this f , we have
kT ∗ f k ≥ |T ∗ f (x)| = |f (T x)| = kT xk > (1 − )kT k. (21.20)
Since was arbitrary, we conclude that kT ∗ k := supkf k=1 kT ∗ f k ≥ kT k.
14
22. Problems
Problem 22.1. a) Prove that if X is a finite-dimensional normed space, then every linear
functional f : X → K is bounded.
b) Prove that if X is any normed vector space, {x1 , . . . xn } is a linearly independent 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.
Problem 22.2. Let X , Y be normed spaces and T : X → Y a linear transformation. Prove
that T is bounded if and only if there exists a constant C such that for all x ∈ X and f ∈ Y ∗ ,
|f (T x)| ≤ Ckf kkxk; (22.1)
in which case kT k is equal to the best possible C in (22.1).
Problem 22.3. Let X be a normed vector space. Show that if M is a closed subspace of
X and x ∈/ M, then M + Cx is closed. Use this to give another proof that every finite-
dimensional subspace of X is closed.
Problem 22.4. Prove that 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 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.)
Problem 22.5. 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 that 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.
Problem 22.6. Prove that if X is a Banach space and X ∗ is separable, then X is separable.
Hint: let {fn } be a countable dense subset of X ∗ . For each n choose xn such that |fn (xn )| ≥
1
2
kfn k. Show that the set of linear combinations of {xn } is dense in X .
Problem 22.7. a) Prove that there exists a bounded linear functional L ∈ (`∞ )∗ with
the following property: whenever f ∈ `∞ and limn→∞ f (n) exists, then L(f ) is equal
to this limit. (Hint: first show that the set of such f forms a closed subspace M ⊂
`∞ ).
b) Show that such a functional L is not equal to Lg for any g ∈ `1 ; thus the map
T : `1 → (`∞ )∗ given by T (g) = Lg is not surjective.
c) Give another proof that T is not surjective, using Problem 22.6.
24. Problems
Problem 24.1. Show that there exists a sequence of open, dense subsets Un ⊂ R such that
m( ∞
T
n=1 n ) = 0.
U
Problem 24.2. Consider the linear subspace D ⊂ c0 defined by
D = {f ∈ c0 : lim |nf (n)| = 0} (24.1)
n→∞
19