Metric Norm
Metric Norm
We collect here miscellaneous problems and results that didn’t make it into
the text, but have some relevance and may be useful. Several notes and
caveats are in order.
1
2 Extra Material and Problems c 2018 Christopher Heil
2.1 Metrics
x · y = x1 y1 + · · · + xd yd , x, y ∈ Fd . (2.15)
and
d∞ (x, y) = max |x1 − y1 |, . . . , |xd − yd | .
Prove that d1 and d∞ are each metrics on Fd . ♦
More generally, we will see in Section 3.2 that if p is any number in the
range 1 ≤ p < ∞, then
1/p
dp (x, y) = |x1 − y1 |p + · · · + |xd − yd |p (2.17)
This function d1/2 does not satisfy the Triangle Inequality, because the choice
x = (1, 0, 0, . . . , 0), y = (1, 1, 0, . . . , 0), and z = (0, 1, 0, . . . , 0) yields
Consequently, equation (2.18) does not define a metric. On the other hand,
the following exercise shows how to create a related but slightly different
function d1/2 that is a metric on Fd .
3.2. Problem 2.1.13 shows that is possible to define a metric on any set X.
However, in practice we usually want to do more than just define an arbitrary
metric on X. Typically, our set X has some kind of interesting properties,
and we seek a metric that somehow takes those properties into account. We
will see many examples of such metrics later. In particular, we will define
metrics on some spaces whose elements are infinite sequences in Section 3.2,
and metrics on some spaces whose elements are functions in Section 3.5.
Extra Problems
2.1.54. (a) Can you define a space of sequences ℓ2 and a corresponding metric
d2 that is analogous to the Euclidean metric on Fd ?
(b) Can you define a metric on C[0, 1] that is analogous in some way to
the Euclidean metric on Fd ?
Extra Problems
Also prove that the same result holds if we replace the Euclidean metric d2
by d1 , d∞ , or d1/2 .
2.2.53. Do you think that the space C[0, 1] is complete with respect to the
uniform metric du or the L1 -metric d1 ? Try to give a variety of different ex-
amples of Cauchy sequences and determine whether they converge in C[0, 1].
We will consider the completeness of C[0, 1] in more detail later, but try to
prove now that it is complete or incomplete with respect to each of these two
metrics.
Neighborhoods
Extra Problems
2.3.51. Let X be a metric space and fix x ∈ X. Prove that if y and z are any
two points in Br (x), then d(x, y) < 2r.
2.3.52. Let X be a metric space, and let x ∈ X and E ⊆ X be fixed. Prove
the following statements.
(a) Every open ball Br (x) is an open subset of X.
(b) A ⊆ X is a neighborhood of x if and only if there exists some r > 0
such that Br (x) ⊆ A.
(c) E is open if and only if E = E ◦ .
(d) E ◦ is the union of all open balls that are contained in E.
2.3.53. Let X be a metric space, and let xn , x ∈ X be given. Prove that the
following three statements are equivalent.
(a) xn → x.
(b) For every open neighborhood U of x, there is an N > 0 such that
xn ∈ U for all n > N.
(c) For every neighborhood A of x, there is an N > 0 such that xn ∈ A
for all n > N.
Extra Problems
is open but not closed, and therefore its complement is closed but not open.
Given an explicit direct description of E C , i.e., f ∈ E C if and only if what?
2.4.52. Let X be a metric set. Given A, B ⊆ X, define
D(A, B) = sup dist(x, B) : x ∈ A + sup dist(y, A) : y ∈ B .
Accumulation Points
xn ∈ Br (x) ⊆ U ⊆ A.
Thus the neighborhood A contains at least one point (and maybe more, if
the xn are not all equal) that belongs to E but is not equal to x. Another
way to say this is that if x is an accumulation point of E, then
E is closed ⇐⇒ E ′ ⊆ E.
A = [0, 1] ∪ {2}
The interval [0, 1] is a perfect set. However, perfect sets can be quite un-
usual. For example, we will see in Problem 2.7.3 that the Cantor set C is
perfect and has no interior. The Cantor set is a closed, uncountable subset
of [0, 1] that contains no intervals, yet every point x ∈ C is an accumulation
point of C!
According to Problem 2.8.59, every perfect subset of Rd is uncountable.
Although we will not prove it, the Cantor–Bendixson Theorem implies that
every nonempty closed set F ⊆ Rd can be uniquely written as F = E ∪ Z
where E is perfect and Z is countable.
Boundary Points
Our next goal is to characterize closed sets in terms of their boundary points.
Here is the definition of an interior point, exterior point, and boundary point
of a set E.
Definition 2.5.54 (Interior, Exterior, and Boundary Points). Let E
be a subset of a metric space X, and let x be any point in X.
(a) We say that x is an interior point of E if there exists a neighborhood A
of x such that A ⊆ E.
(b) We say that x is an exterior point of E if there exists a neighborhood A
of x such that A ⊆ E C .
(c) We say that x is a boundary point of E if for every neighborhood A of x
we have both A ∩ E 6= ∅ and A ∩ E C 6= ∅. The set of all boundary points
of E is called the boundary of E, and it is denoted by
∂E = x ∈ X : x is a boundary point of E . ♦
Boundary points and accumulation points are similar in some respects, but
they are not identical. For example, 0 is a boundary point of the singleton {0},
but it is not an accumulation point. This shows that boundary points need
not be accumulation points. On the other hand, only 0 and 1 are boundary
points of the interval [0, 1], yet every point in [0, 1] is an accumulation point.
Therefore accumulation points need not be a boundary points.
According to Theorem 2.5.5, a subset of a metric space is closed if and
only if it contains all of its boundary points, i.e., ∂E ⊆ E.
Can we ever have ∂E = E? If L is the line segment
L = {(x, x) : 0 ≤ x ≤ 1}
∂L = {a, b} =
6 [a, b] = L.
10 Extra Material and Problems c 2018 Christopher Heil
It is hard to imagine an infinite subset of the real line that equals its own
boundary! Yet Problem 2.7.3 shows that the Cantor set is an example of an
uncountable subset of the real line that has this unusual property.
Thus C is a rather strange set. It is uncountable, yet its interior is empty.
It equals its own boundary, and every point in C is both an accumulation
point of C and an accumulation point of its complement R \ C. Following
Definition 2.5.53, since every point in C is an accumulation point of C, we
say that C is a perfect set. It is also totally disconnected, which means that
it contains no nontrivial connected subsets (in one dimension, a nontrivial
connected set is simply an interval).
d(x, y) = |x − y|, x, y ∈ Q.
Let E be the set of all rational numbers that are less than π, i.e.,
Let x1 = 3.1, x2 = 3.14, x3 = 3.141, x4 = 3.1415, and so forth. Then {xn }n∈N
is a Cauchy sequence in Q (why?), but it does not converge to an element
of Q. Therefore E is not a complete subset of Q.
On the other hand, we will show that E is a closed subset of Q. Suppose
that {yn }n∈N is any sequence of points from E that converges to some point
y ∈ Q. By definition of E, we have yn < π for every n. Since yn converges
to y, it follows that
y = lim yn ≤ π.
n→∞
Thus, a closed set need not be a complete set in general. Problem 2.4.8
shows that if X itself is complete then the closed subsets of X coincide with
the complete subsets of X.
2.5 Accumulation Points and Boundary Points 11
Extra Problems
Density
Observe that Theorem 2.6.5 implies that the Banach space ℓ1 contains proper,
dense subspaces. This is quite different from our experience with the Eu-
clidean space Rd , whose proper subspaces are closed sets and therefore are
not dense in Rd . Indeed, Problem 3.7.5 shows that no proper subspace of any
finite-dimensional normed space can be dense.
Separability
and we can find uncountably many vectors in ℓ∞ that are each a distance 1
from each other!
Problem 2.6.20 shows that ℓ1 and c0 are separable, and Theorem 2.6.8
shows that ℓ∞ is not separable. his may seem to be related to the fact that
ℓ 1 ⊆ c0 ⊆ ℓ ∞ ,
but this is misleading, because these spaces do not all share the same norms.
Indeed, ℓ1 is separable with respect to the ℓ1 -norm, c0 is separable with
respect to the ℓ∞ -norm, and ℓ∞ is not separable with respect to the ℓ∞ -norm.
Problem 2.6.22 does show that if Y is a subspace of a separable space X and
we use the norm on Y inherited from X, then Y must be separable. However,
if Y is given a different norm, then it is entirely possible that Y could be
nonseparable even though X is separable. Although we will not prove it here,
an example from measure theory is given by the Lebesgue spaces L1 [0, 1] and
L∞ [0, 1]. The space L∞ [0, 1] is nonseparable, even though it is a proper subset
of the separable space L1 [0, 1] (for precise definitions and proof, see [Heil18]).
Again, the norms on these two spaces are different. If we replace the standard
norm on L∞ [0, 1] with the norm of L1 [0, 1], then L∞ [0, 1] is a separable space.
Extra Problems
2.6.52. Let
A = x = (xk )k∈N ∈ ℓ1 : x1 = 0 .
(a) Prove that A is not a dense subset of ℓ1 by showing that there exists
some point z ∈ ℓ1 such that there do not exist any points xn ∈ A that
converge to z in ℓ1 -norm.
(b) Prove that A is a proper, closed subspace of ℓ1 , and use this to give
another proof that A is not dense in ℓ1 .
Extra Problems
Show that D + D = [0, 1], and use this to show that C + C = [0, 2].
We know that a closed set contains all of its accumulation points. In fact,
Theorem 2.4.2 tells us that if {xn }n∈N is any convergent sequence that is
contained in a closed set F, then the limit of this sequence must belong
to F. Even so, it is not true that every infinite sequence {xn }n∈N in F must
converge, or even have an accumulation point or a convergent subsequence.
For example, the set of integers Z is a closed subset of the real line R, but if we
set xn = n then {xn }n∈N does not converge, and no subsequence of {xn }n∈N
converges. One of the equivalences proved in Theorem 2.8.9 shows that any
infinite sequence in a compact set must contain a convergent subsequence.
We will use Theorem 2.8.9 to give an example of a closed and bounded set
that is compact.
Proof. Implicitly, we are assuming that the metric on R is its standard metric
(the absolute value metric).
The interval [0, 1] is a closed subset of R. Since R is complete, Problem
2.4.8 implies that [0, 1] is complete.
Fix any r > 0, and let n be an integer larger than 1r . For each k = 0, . . . , n,
let
Bk = nk − 2r , nk + 2r .
1
Then Bk is an open interval of radius r. Since r > n, it follows that
S
n
[0, 1] ⊆ Uk .
k=0
2.8 Compact Sets in Metric Spaces 15
Thus [0, 1] is covered by finitely many open balls of radius r. Since we can do
this for every r > 0, we conclude that [0, 1] is totally bounded.
Thus [0, 1] is both complete and totally bounded, so Theorem 2.8.9 implies
that [0, 1] is a compact subset of R. ⊓ ⊔
The proof of Theorem 2.8.4 shows that a subset of Rd is compact if and only
if it is closed and bounded. We could use essentially the same proof to show
that a subset of Cd is compact if and only if it is closed and bounded. Instead,
we will indicate a different (and possibly more difficult) approach based on
the relationship between the topologies of R2d and Cd . Our intuition suggests
that C “looks just like” R2 , and indeed the following exercise will show that
there is a natural identification between the open subsets of Cd and the open
subsets of R2d . We assume in this exercise that the norms on Cd and R2d are
the Euclidean norms.
and
Ve = (a, b) ∈ R2 : a + ib ∈ V is an open subset of R2 .
Formulate and prove a corresponding result for Cd and R2d . ♦
is a compact subset of R2d and is therefore closed and bounded in R2d . Use
this to prove that K is a closed and bounded set in Cd . ♦
Extra Problems
2.8.55. For each of the following statements, exhibit a metric space X that
has that property.
(a) X is compact.
(b) X is not compact.
(c) X is infinite, but no infinite subset of X is compact.
(d) X is infinite but not compact, yet there exists an infinite subset of X
that is compact.
Continuity at a Point
(c) For each open neighborhood V of f (x) in Y, there exists an open neigh-
borhood U of x in X such that U ⊆ f −1 (V ). ♦
Proof. (a) ⇒ (b). Assume that f is continuous at x, and suppose that xn → x.
Given ε > 0, there is a δ > 0 such that equation (2.20) holds. Also, by
definition of convergent sequences, since xn → x there is an N > 0 such that
dX (xn , x) < δ for all n > N. Therefore dY (f (xn ), f (x)) < ε for all n > N, so
we have f (xn ) → f (x) in Y.
(b) ⇒ (c). Suppose that statement (c) fails, i.e., there exists some open
neighborhood V of f (x) such that no open neighborhood U of x is contained
in f −1 (V ). Since U = B1/n (x) is an open neighborhood of x, there must be
a point xn ∈ B1/n (x) that is not in f −1 (V ). Hence xn → x, but f (xn ) ∈
/V
for any n so f (xn ) cannot converge to f (x). Therefore statement (b) fails.
(c) ⇒ (a). Suppose that statement (c) holds, and fix any ε > 0. Then
V = Bε (f (x)) is an open neighborhood of f (x) in Y, so statement (c) implies
that there exists an open neighborhood U of x in X such that U ⊆ f −1 (V ).
Since x ∈ U and U is open, there is a δ > 0 such that Bδ (x) ⊆ U. If y is any
point in X that satisfies dX (x, y) < δ, then we have
y ∈ Bδ (x) ⊆ U ⊆ f −1 (V ).
xn → x in X =⇒ f (xn ) → f (x) in Y. ♦
Example 2.9.53. We know from undergraduate calculus class that the func-
tion f : R → R defined by f (x) = sin x is continuous. If V is any open subset
of the real line, then its inverse image f −1 (V ) is open. For example, the
inverse image of the interval V = (0, 2) is
which is open. However, the direct image of the open interval U = (0, 2π) is
f (U ) = [−1, 1],
Examples
These examples use some function spaces that are introduced in Chapter 3.
We will illustrate continuity by considering a function that maps the
normed space Cb1 (R) into the space Cb (R). The space Cb (R) consists of all
bounded, continuous functions, while Cb1 (R) is the space of all bounded, dif-
ferentiable functions f such that f ′ is bounded and continuous. As discussed
in Section 3.5, the standard norm for Cb (R) is the uniform norm, while the
standard norm for Cb1 (R) is kf kCb1 = kf ku + kf ′ ku . Unless we specify other-
wise, we always assume that these are the norms on these two spaces.
Example 2.9.54. Given g ∈ Cb1 (R), let Dg = g ′ be the derivative of g. By
definition of Cb1 (R), the derivative g ′ belongs to Cb (R). Therefore we can think
of D as a function whose domain is Cb1 (R) and whose codomain is Cb (R). We
input a function g from Cb1 (R), and we output the function Dg = g ′ ∈ Cb (R).
Normally we would write D(g) for the output of the function D, but in this
context it is traditional, and simpler, to just write Dg.
We claim that D : Cb1 (R) → Cb (R) is continuous. Note that the elements of
Cb (R) are functions. Hence a “point” in Cb1 (R) is a function g. The function
1
This is easy! To see why, suppose that gn → g in Cb1 (R). This means that gn
converges to g with respect to the norm of Cb1 (R), or in other words
lim kg − gn kCb1 = 0.
n→∞
20 Extra Material and Problems c 2018 Christopher Heil
Thus Dgn converges to Dg with respect to the uniform norm, which is the
norm of Cb (R). In summary, we have shown that if gn → g in Cb1 (R), then
Dgn → Dg in Cb (R). Therefore D is continuous at the point g. Since g is an
(
arbitrary function in Cb R), we conclude that D is continuous on the domain
Cb1 (R). ♦
Example 2.9.55. Let us point out a few other properties of the derivative
function D considered in Example 2.9.54.
First, D is not injective, because we can find two different functions f, g in
the domain Cb1 (R) such that Df = Dg. In fact, if f is any function in Cb1 (R)
and c is a constant, then g = f + c has the same derivative as f.
Second, D is not surjective, because it is not true that the range of D
equals the codomain Cb (R). In particular, if g is the constant function 1,
then there is no function f ∈ Cb1 (R) that satisfies Df = 1. In fact, the only
functions whose derivative is the constant 1 are f (x) = x + c where c is
a constant, but none of these functions belong to Cb1 (R) because they are
unbounded.
Third, D is a linear function, because if f, g belong to Cb1 (R) and a, b are
scalars, then
Uniform Continuity
Example 2.9.56. Let D : Cb1 (R) → Cb (R) be the derivative operator discussed
in Example 2.9.55. We will prove that D is uniformly continuous. Given ε > 0,
let δ = ε. Our “points” x, y from Definition 2.9.5 are now elements of Cb1 (R),
2.9 Continuity for Functions on Metric Spaces 21
i.e., they are functions in Cb1 (R). So, suppose that f, g are any two elements
of Cb1 (R) such that kf − gku < δ. Then
kDf − Dgku = kf ′ − g ′ ku
≤ kf − gku + kf ′ − g ′ ku
= kf − gkCb1
< δ = ε.
Thus kDf − Dgku < ε simultaneously for all f and g in Cb1 (R). This shows
that D is uniformly continuous on Cb1 (R). ♦
Extra Problems
defines a metric on X × Y.
(b) Suppose that f : X → Y is continuous, and prove that
graph(f ) = (x, f (x)) : x ∈ X
is a closed subset of X × Y.
(c) Suppose that X is compact. Prove that f : X → Y is continuous if and
only if
2.9 Continuity for Functions on Metric Spaces 23
graph(f ) = (x, f (x)) : x ∈ X
is a compact subset of X × Y.
The simplest vector space, aside from the trivial space X = {0}, is the field
of scalars F. The “standard” norm on F is absolute value. However, this is
not the only norm, for if λ > 0 is a fixed positive real number then
kxk = λ |x|, x ∈ F,
defines another norm on F. Are there any other norms on F? Yes, but we
can identify all of them—according to Problem 3.1.8, a function k · k is a
seminorm on F if and only if there exists a scalar λ ≥ 0 such that kxk = λ |x|
for every x ∈ F; furthermore, this seminorm is a norm if and only if λ > 0.
We often summarize this by saying that up to multiplication by a positive
scalar, absolute value is the only norm on F, and the only seminorm on F
that is not a norm is the zero function.
The situation in higher dimensions is more complicated. One seminorm on
Fd is the zero function, defined by kxk = 0 for all x ∈ Fd . However, when
d ≥ 2 there are many nonzero functions on Fd that are seminorms, such as
Extra Problems
3.1.51. (a) Fix a dimension d ≥ 1. Show that there exist positive constants
Ad , Bd such that
The constants Ad , Bd can depend on the dimension d, but not on x. Find the
optimal constants Ad , Bd , i.e., find the largest constant Ad and the smallest
constant Bd such that equation (3.23) holds for every x ∈ Fd .
(b) Repeat part (a), but with k · k2 in place of k · k∞ .
(c) Repeat part (a), but with k · k2 in place of k · k1 .
Remark: Using the terminology of Definition 3.6.1, this problem says that
k · k1 , k · k2 , and k · k∞ are all equivalent norms on Fd .
Using this notation, the symbol xk denotes the kth component of x. However,
it is sometimes preferable to write
x = x(k) k∈N = x(1), x(2), . . . .
That is, it is sometimes more clear to let x(k) denote the kth component
of x. In any particular situation we will use whichever of these two notations
is more convenient.
Remark 3.2.50. By making appropriate changes in the definitions above, we
can consider spaces of sequences that are indexed by sets other than the
natural numbers N. For example, if I is a countable index set, then we say
that a sequence x = (xk )k∈I is p-summable if and only if
X
|xk |p < ∞.
k∈I
We let ℓp (I) be the space of all p-summable sequences indexed by I, and let
ℓ∞ (I) be the space of all bounded sequences indexed by I. If I = N, then
this reduces to the definition of ℓp that we gave before.
A common choice of index set is I = Z. A sequence indexed by Z is a
bi-infinite sequence of the form
26 Extra Material and Problems c 2018 Christopher Heil
Hence ℓp (Z) denotes the set of all bi-infinite sequences that are p-summable
(if p is finite) or bounded (if p = ∞). For example, the bi-infinite sequence
x = 2−|k| k∈Z = . . . , 41 , 12 , 1, 21 , 14 , . . .
and therefore x + y ∈ ℓp .
= 2p max{|a|p , |b|p }
≤ 2p |a|p + |b|p .
∞
X ∞
X
kx + ykpp = |xk + yk |p ≤ 2p |xk |p + |yk |p = 2p kxkpp + kykpp .
k=1 k=1
0 = (0, 0, 0, . . . ).
Let
δn = (0, . . . , 0, 1, 0, 0, . . . )
denote the sequence that has a 1 in the nth component and zeros elsewhere.
We call δn the nth standard basis vector, and refer to E = {δn }n∈N as the
sequence of standard basis vectors, or simply the standard basis.
Note that E = {δn }n∈N is a sequence, and each element δn of this sequence
is itself a sequence of scalars. Thus E is a “sequence of sequences,” but that
is not necessarily the best way to think about E. Instead, it is better to think
of E as being a sequence of vectors, where it just so happens that “vector”
here means a sequence of scalars. Each δn is a vector in ℓp , and E is just the
sequence of vectors E = {δ1 , δ2 , . . . } in ℓp .
28 Extra Material and Problems c 2018 Christopher Heil
(c1 , . . . , cN , 0, 0, . . . ) = (0, 0, 0, . . . ).
E. For example,
z = 2−n n∈N
= 1 1 1
2, 4, 8, . . .
span(E) ( ℓp , 0 < p ≤ ∞.
Indeed, span(E) is precisely the set of all sequences that have finitely many
nonzero components. We will study this space, which we often denote by
c00 = span(E), in more detail in Example 2.2.13.
The fact that ℓp contains an infinite linearly independent set implies that
it must be infinite dimensional. This is because a basic result from linear
algebra tells us that if a vector space X has finite dimension d, then any set
of more than d vectors must be dependent. A finite-dimensional vector space
cannot contain an infinite linearly independent set.
Hölder’s Inequality
Fig. 3.50 The curved line is the graph of y = xp−1 . The area of the vertically hatched
1
region is 0a xp−1 dx, the area of the horizontally hatched region is 0b y p−1 dy, and the
R R
The method of Problem 3.2.11 is not the only way to prove equation (3.8),
which stated that if 1 < p < ∞ then
′
ap bp
ab ≤ + ′, a, b ≥ 0.
p p
30 Extra Material and Problems c 2018 Christopher Heil
For another approach, note that xp−1 is continuous and strictly increasing on
1
the interval [0, a], and its inverse function is y p−1 . Figure 3.50 gives a “proof
by picture” that
Z a Z b
1
ab ≤ xp−1 dx + y p−1 dy. (3.27)
0 0
Extra Problems
3.2.53. Given 0 < p < 1, prove that ℓp is complete with respect to the metric
defined in Problem 3.3.14.
Convexity
If a metric space contains an open ball that is not convex, then there is an
important further conclusion that we can draw. This is spelled out in the
following lemma.
Lemma 3.3.50. Let X be a vector space that has a metric d. If there exists an
open ball Br (x) in X that is not convex, then the metric d is not induced from
any norm on X. That is, there is no norm k·k on X such that d(y, z) = ky−zk
for all y, z ∈ X.
Proof. Suppose that the open ball Br (x) is convex. Suppose also that there
is a k · k on X that satisfies ky − zk = d(y, z) for all y, z ∈ X. By Problem
3.3 The Induced Metric 31
3.3.13, open balls defined with respect to a norm are convex. Yet the open
ball centered at x with radius r that is defined with respect to k ·k is precisely
Br (x):
Br (x) = y ∈ X : d(y, x) < 1 = y ∈ X : ky − xk < 1 .
Bounded Sets
Separability
Suppose now that U is any set that is open with respect to the Euclidean
norm. By definition, this means that for each x ∈ U there exists some rx > 0
such that Br2x(x) ⊆ U. Hence,
S S
U = {x} ⊆ Br1x(x) (since x ∈ Br1x(x))
x∈U x∈U
S
⊆ Br2x(x) (by equation (3.28))
x∈U
This implies that U is the union of the open balls Br1x(x) over x ∈ U, i.e.,
S
U = Br1x(x).
x∈U
Since each such ball Br1x(x) is open with respect to k · k1 , it follows that U is
open with respect to k · k1 .
If we attempt to use the same argument to prove the converse implication,
then we encounter a problem because Br2 (x) is not contained in Br1 (x). On
the other hand, by applying Hölder’s Inequality with the index p = 2 we see
that
d
X X
d 1/2 X
d 1/2
kxk1 = |1 · xk | ≤ 12 |xk |2 = d1/2 kxk2 ,
k=1 k=1 k=1
That is, every open disk is contained in a diamond whose radius has been
enlarged by a factor of d1/2 . We can then argue just as before to show that
if U is open with respect to the norm k · k1 , then it is also open with respect
to k · k2 . ⊓
⊔
Extra Problems
3.3.54. Show that if X is a nontrivial normed space, then every open ball
Br (x) is an infinite, proper subset of X, and r>0 Br (x) = X.
S
Complete Subsets
Remark 3.4.51. As we have noted before, the term “complete” has a num-
ber of distinct mathematical meanings. In this volume, the two main uses
are complete sets in the sense of Definition 2.2.9, and complete sequences in
the sense of Definition 4.4.4. The reader should pause to consider context
whenever the term “complete” is encountered. ♦
Example 3.4.52. Let S be the open interval S = (0, 1) in the real line. Then
{ n1 }n∈N is a Cauchy sequence of elements of S, but this sequence does not
converge to an element of S. The sequence does converge, but it does not
converge to a point in S. Therefore S is not a complete subset of R. ♦
Theorem 3.4.53. The sup-norm is a norm on c00 , but c00 is not complete
with respect to k · k∞ .
and consider the sequence of vectors {xn }n∈N in c00 . If m < n, then
1
xn − xm = 0, . . . , 0, m+1 , . . . , n1 , 0, 0, . . . ,
so
1
kxn − xm k∞ = .
m+1
Therefore, if we fix ε > 0 then we have kxn − xm k∞ < ε for all m, n > 1ε .
Hence {xn }n∈N is a Cauchy sequence in c00 .
Suppose that there was a sequence x ∈ c00 such that xn → x with respect
to the sup-norm. We know that ℓ∞ -norm convergence implies componentwise
convergence, so xn must converge componentwise to x as n → ∞. If we fix
any particular k, then
1
xn (k) = for all n > k.
k
Therefore we must have x(k) = k1 . That is, x is the sequence
x = 1, 21 , 13 , . . . . (3.29)
Extra Problems
Remark 3.5.50. The uniform norm is the analogue for functions of the sup-
norm
kxk∞ = k(xk )k∈N k∞ = sup |xk | < ∞
k∈N
that is defined for sequences. It would therefore be natural to use the notation
kf k∞ instead of kf ku to denote the uniform norm of f. However, the symbols
kf k∞ traditionally denote the L∞ -norm of a function f. If f is a continuous
function on an interval I, then there is no difference between its L∞ -norm
and its uniform norm. However, for discontinuous functions there can be a
difference, because the L∞ -norm “ignores sets of measure zero” (in a sense
that is made precise by using measure theory; see [Heil18]). For this reason,
we will always write kf ku to denote the uniform norm of a function. ♦
3.5 Uniform Convergence of Functions 37
Now we look at continuous functions whose domain is the entire real line R.
We will define two important subspaces of Cb (R). To define the first of these,
we declare that a function f : R → F vanishes at infinity if we have both
belongs to C0 (R). Since φ(x) 6= 0 for any x, this shows that an element
of C0 (R) need not be zero at any point.
(b) The function f (x) = e−x does not belong to C0 (R) because f (x) does not
converge to zero as x → −∞.
(c) The two-sided exponential g(x) = e−|x| is an element of C0 (R), and it
satisfies g(x) 6= 0 for every x.
(d) The sinc function is
sin πx
sinc(x) = , x 6= 0.
πx
The sinc function has a removable discontinuity at the origin, so if we
define sinc(0) = 1 then it belongs to C0 (R). The sinc function has the
property that sinc(x) = 0 if and only if x is a nonzero integer (this plays
an important role in the Classical (or Shannon) Sampling Theorem; see
[Heil11]). ♦
Each function in Cc (R) belongs to C0 (R), and Cc (R) is closed under addi-
tion and multiplication by scalars, so it is a subspace of C0 (R). Each of the
functions in parts (a), (c), and (d) of Example 3.5.51 belong to C0 (R) but
not Cc (R). Hence Cc (R) is a proper subspace of C0 (R). More generally, we
have the inclusions
Example 3.5.54. Suppose that instead of Cc (R), we consider the set S that
consists of those continuous, compactly supported functions that are identi-
cally zero outside of the interval [0, 1], i.e.,
S = g ∈ Cc (R) : supp(g) ⊆ [0, 1] .
In particular, if we choose x outside of [0, 1], then fn (x) = 0 for every n and
therefore f (x) = 0 as well. Hence f is compactly supported and is identically
zero outside of [0, 1], so f belongs to S. Applying Theorem 2.4.2, we conclude
that S is a closed subset of C0 (R). ♦
1 1
0.75 0.75
j1 j2
0.5 0.5
0.25 0.25
0 0
0 0.25 0.5 0.75 1 0 0.25 0.5 0.75 1
0.75
j3
0.5
0.25
0
0 0.25 0.5 0.75 1
the construction of the Cantor set, and it is linear on the remaining subin-
tervals of [0, 1]. The function ϕ2 takes the same constant 21 on the interval
( 13 , 32 ) but additionally is constant with values 41 and 43 on the two intervals
that are removed during the second stage of the construction of the Cantor
set. We continue this process and define ϕ3 , ϕ4 , . . . in a similar fashion. Each
function ϕn is continuous, and ϕn is constant on each of the open intervals
that are removed during the nth stage of the construction of the Cantor set.
1 Looking
2
at the graphs of ϕ1 and 1ϕ2 , observe
2 that ϕ1 = ϕ2 on the interval
,
3 3 , and if x belongs to either 0, 3 or 3 , 1 then |ϕ1 (x)−ϕ2 (x)| is at most
1 1
2 . Hence kϕ 1 − ϕ k
2 u ≤ 2 . Continue this argument, and prove the following
statements.
(a) For each n ∈ N, ϕn is continuous and monotone increasing on [0, 1], i.e.,
if 0 ≤ x ≤ y ≤ 1, then ϕn (x) ≤ ϕn (y).
(b) For each n ∈ N,
(c) {ϕn }n∈N is a Cauchy sequence in C[0, 1] with respect to the uniform
norm.
3.5 Uniform Convergence of Functions 41
We know that C[0, 1] is a Banach space with respect to the uniform norm.
Therefore, since {ϕn }n∈N is a Cauchy sequence in C[0, 1], there must exist a
function ϕ ∈ C[0, 1] such that ϕn converge uniformly to ϕ as n → ∞! Since
uniform convergence implies pointwise convergence, we have
1 2
Fig. 3.52 The reflected Devil’s staircase (Cantor–Lebesgue function).
Recall that given an interval I in the real line, we let Cb (I) be the set of all
functions f : I → F that are both bounded and continuous. The space Cb1 (I)
consists of all differentiable functions f : I → F such that both f and f ′ are
bounded (we focused on Cb1 (R) in Section 3.50, but the same ideas apply to
functions on an interval I). By definition, we have
K = kf ′ ku = sup |f ′ (t)|
t∈I
is finite. Choose any two points x < y in I. Because f is real-valued, the Mean
Value Theorem (Theorem 1.9.1) implies that there exists a point c between
x and y such that
f (y) − f (x)
f ′ (c) = .
y−x
Rearranging, we see that
Hence the distance between f (x) and f (y) is never more than K times the
distance between x and y. Such a function does not expand the distance be-
tween points by more than the factor K (technically the word “compression”
may be more appropriate if K < 1, but the words “expand” or “dilate” are
often used in this context regardless of whether K is larger or smaller than 1).
Every real-valued differentiable function whose derivative is bounded has this
3.5 Uniform Convergence of Functions 43
property of limited expansion. On the other hand, the following exercise shows
that there exist non-differentiable functions that have the same property.
Exercise 3.5.56. The hat function W (x) = max 1 − |x|, 0 is piecewise
linear. Prove that W satisfies
Any number K such that equation (3.32) holds is called a Lipschitz constant
for f. ♦
The Lipschitz constant is not unique, for if K is a Lipschitz constant for f
then so is any number larger than K. The smallest, or “optimal,” Lipschitz
constant for f is
|f (y) − f (x)|
K = sup .
x6=y |x − y|
If f is Lipschitz and we choose any ε > 0, then |f (x) − f (y)| < ε whenever
|x − y| < δ = ε/K. Hence every Lipschitz constant is uniformly continuous.
We collect the bounded Lipschitz functions to form the following space.
Definition 3.5.58. Given an interval I, we set
Lip(I) = f ∈ Cb (I) : f is Lipschitz continuous on I . ♦
As a corollary, we see that Cb1 (I) ⊆ Lip(I). The hat function W is Lipschitz
but not differentiable on the interval [−1, 1]. If I is an arbitrary interval, then
by appropriately dilating and translating the function so that it is supported
within I, we obtain a function that is Lipschitz on I but does not belong to
Cb1 (I). Hence Cb1 (I) is a proper subspace of Lip(I). Another very interesting
example is constructed in part (c) of Problem 3.5.74, which shows that there
exists a function that is Lipschitz and differentiable everywhere on I but does
not belong to Cb1 (I) because its derivative is not continuous.
By definition, Lip(I) ⊆ Cb (I). Are there bounded continuous functions
that are not Lipschitz? Yes; here is an example.
Example 3.5.60. The Cantor–Lebesgue function ϕ takes the value 2−k at the
point x = 3−k . Therefore,
= ∞.
In this sense, the space of Lipschitz functions is intermediate between Cb1 (I)
and Cb (I).
Problem 3.5.76 shows that although the Cantor–Lebesgue function is not
Lipschitz, it does satisfy a weaker Hölder continuity condition. Problem 4.2.15
shows that the space of Lipschitz functions is a Banach space with respect
to an appropriate norm, and Problem 3.5.77 does the same for the space of
Hölder continuous functions.
3.5 Uniform Convergence of Functions 45
Extra Problems
3.5.61. Show that if I is an interval in the real line other than [a, b], then
Cb (I) ( C(I).
3.5.62. Give an example of a Banach space X that contains a dense sub-
space S and a closed subspace M such that S ∩ M = {0}.
3.5.63. (a) Let X, Y be metric spaces. Given a function f : X → Y, prove
that f is continuous if and only if the following implication holds:
(b) Show by example that the direct image of a closed set under a contin-
uous function need not be closed.
(c) Show by example that the inverse image of a compact set under a
continuous function need not be compact.
3.5.64. We say that a function f : R → F is periodic if there exists some
a ∈ R such that f (x − a) = f (x) for every x ∈ R. Prove that every periodic
function f ∈ C(R) is bounded and uniformly continuous on R.
3.5.65. Let X be a metric space, and suppose fn , f ∈ Cb (X). Show that if
fn → f uniformly and xn → x in X, then fn (xn ) → f (x).
3.5.66. (a) Suppose that fn , f ∈ Cb (R), and fn converges pointwise to f,
i.e., fn (x) → f (x) for each x ∈ R. Show that
def
kf ku ≤ lim inf kfn ku = sup inf kfk ku .
n→∞ n∈N k≥n
3.5.70. Recall that C(R) is the space of all continuous scalar-valued functions
on R. This space includes unbounded functions (such as f (x) = x2 ), so the
uniform norm k · ku is not a norm on C(R) (why not?).
For each N ∈ N, define
3.5.71. Given a metric space X, let C(X) be the space of all continuous,
scalar-valued functions f : X → F, and let Cb (X) be the subspace of C(X)
that contains all bounded, continuous functions f : X → F.
(a) Prove that Cb (X) is a vector space, the uniform norm
is a norm on Cb (X), and Cb (X) is a Banach space with respect to the uniform
norm.
(b) Prove that Cb (X) is closed with respect to products of functions, i.e.,
if f, g ∈ Cb (X) then f g ∈ Cb (X). Further, show that the following submulti-
plicative norm inequality is satisfied:
(c) Suppose that g ∈ Cb (X) and there exists an ε > 0 such that |g(x)| ≥ ε
for every x ∈ X. Prove that 1/g ∈ Cb (X).
(d) Show that if X is compact, then
3.5 Uniform Convergence of Functions 47
C(X) = Cb (X)
For example, if f (x) = cos x, then δ(f ) = cos 0 = 1. Prove that δ is continu-
ous. Does δ belong to Cb (X)?
n
gn
1
0 n
1
(b) Given an interval [c, d] ⊆ [0, 1], let W[c,d] be the hat function of height 1
on the interval [c, d]. Explicitly,
0, a ≤ x ≤ c,
c+d
linear, c < x < 2 ,
W[c,d] = 1, c+d
x= 2 ,
linear, c+d
2 < x < d,
0, d ≤ x ≤ b.
Define
h1 = W[0,1] ,
h2 = W[0, 12 ] , h3 = W[ 12 ,1] ,
h4 = W[0, 31 ] , h5 = W[ 13 , 32 ] , h6 = W[ 32 ,1] ,
48 Extra Material and Problems c 2018 Christopher Heil
and so forth. Picturing the graphs of these functions as triangles, the triangles
march from left to right across the interval [0, 1], then shrink in size and march
across the interval again, and do this over and over (see Figure 3.54).
1
h7 h8 h9 h10
0.5
0
0 0.25 0.5 0.75 1
Fig. 3.54 Graphs of the functions h7 , h8 , h9 , and h10 from Problem 3.5.72(b).
Prove that T satisfies the hypotheses of the Banach Fixed Point Theorem (see
Problem 2.9.20), and the fixed point of T satisfies the differential equation
f ′ (x) = xf (x) + 1.
/ C 1 [−1, 1].
(c) If a = 1 + b, then f ∈
(d) Show that the function f defined in Example 3.50.50 is differentiable at
every point, f ′ is bounded but not continuous on [−1, 1], and f ∈ Lip[−1, 1].
3.5.77. Given 0 < α < 1, let C α (I) be the space of all bounded functions
that are Hölder continuous with exponent α on I, i.e.,
C α (I) = f ∈ Cb (I) : f is Hölder continuous with exponent α .
|f (x) − f (y)|
kf kC α = kf ku + sup , f ∈ C α (I).
x6=y |x − y|α
ΦHyL
tΦHxL+H1-tLΦHyL
ΦHxL
a x tx+H1-tLy y b
In other words, as illustrated in Figure 3.55, on any subinterval [x, y] of (a, b),
the graph of φ lies on or below the line segment that joints the points (x, φ(x))
and (y, φ(y)). An analogous definition is made for concave functions.
(a) Assume φ : (a, b) → R is a convex function. Prove the Discrete Jensen
Inequality: If N ≥ 2, then for any points x1 , . . . , xN ∈ (a, b) and positive
weights t1 , . . . , tN that satisfy t1 + · · · + tN = 1, we have
X
N N
X
φ tj xj ≤ tj φ(xj ).
j=1 j=1
(b) Given a convex function φ on (a, b) and given x ∈ (a, b), prove that
the function
φ(y) − φ(x)
m(y) = , y ∈ (a, b), y 6= x,
y−x
is monotone increasing on (a, x) and monotone increasing on (x, b). Also show
that if z < x < y, then m(z) ≤ m(y).
(c) Let φ : (a, b) → R be given. Prove that φ is convex if and only if for all
a < x < y < z < b we have
(g) Use the Discrete Jensen Inequality and the fact that ex is convex to
give another proof of the fact that if a, b ≥ 0 and 1 < p < ∞ then
′
ap bp
ab ≤ + ′.
p p
3.5.79. Fix −∞ ≤ a < b ≤ ∞, and assume that φ is a convex function on
(a, b). Prove that φ has the following properties.
(a) φ is right-differentiable on (a, b), i.e., for each point x ∈ (a, b) the limit
φ(y) − φ(x)
φ′+ (x) = lim
y→x+ y−x
φ(y) − φ(x)
φ′+ (x) ≤ ≤ φ′− (y),
y−x
(e) Suppose that y is a point of continuity for φ′+ . Use part (b) to show
that
φ(y) − φ(x)
φ′− (y) = lim− ≥ lim− φ′+ (x) = φ′+ (y).
x→y y−x x→y
Extra Problems
(a) k · ka ≍ k · kb .
(b) There exists some C > 0 such that
1
kxka ≤ kxkb ≤ C kxka , x ∈ X.
C
(c) There exists some r > 0 such that
a
B1/r (0) ⊆ B1b (0) ⊆ Bra (0).
(u − v)
y = .
ku − vk
This extra section expands on the material in Section 3.5 where the space
Cb1 (I) is introduced and discussed.
Most of Section 3.5 deals with spaces whose elements are continuous func-
tions. We will generalize some of the results obtained there to spaces of m-
times differentiable functions. For simplicity of presentation, in this section
we we will take the domain of our functions to be I = R. However, only
small modifications need to be made if we want to consider functions whose
domain is a different type interval (essentially, one-sided instead of two-sided
limits need to be used at any boundary point of the interval).
We begin with a technical issue that complicates the issue of how best to
define spaces of differentiable functions. The problem is that it is not true
that a differentiable function must have a continuous derivative. Here is an
example of a function that is differentiable at every point, yet its derivative
is not continuous.
0.2
0.1
-0.2
1.0
0.5
Fig. 3.56 The function f from Example 3.50.50 (top), and its derivative f ′ (bottom), for
0 < x < 0.4. Note the differing vertical scales on the two graphs.
This function is continuous at every point (why?). We will prove that f ′ (x)
exists at every point x, yet f ′ is discontinuous at x = 0.
To see that f is differentiable at points other than x = 0, note that the
function 1/x is differentiable on (0, ∞) and sin x is differentiable on all of R.
56 Extra Material and Problems c 2018 Christopher Heil
f (0 + t) − f (0) t2 sin(1/t)
f ′ (0) = lim = lim = lim t sin(1/t) = 0.
t→0 t−0 t→0 t t→0
For this and other reasons, the space of all differentiable functions can be
difficult to work with. Usually we impose some restrictions to create a “nicer”
space. For example, we can consider the space of all differentiable functions
which have a continuous derivative. We call this space C 1 (R). Extending to
higher orders, we let C m (R) denote the set of all of all m-times differentiable
functions on R such that f and each derivative f ′ , . . . , f (m) is continuous,
i.e.,
C m (R) = f ∈ C(R) : f, f ′ , . . . , f (m) ∈ C(R) .
On the preceding line, when we write f, f ′ , . . . , f (m) ∈ C(R) we implicitly
mean that f, f ′ , . . . , f (m) all exist (i.e., f is m-times differentiable) and each
of these functions belongs to C(R).
The space C m (R) is still too “large” to define a convenient norm, so we
look at the following subspaces of C m (R):
Cbm (R) = f ∈ Cb (R) : f, f ′ , . . . , f (m) ∈ Cb (R) ,
C0m (R) = f ∈ C0 (R) : f, f ′ , . . . , f (m) ∈ C0 (R) ,
Ccm (R) = f ∈ Cc (R) : f, f ′ , . . . , f (m) ∈ Cc (R) .
3.50 Extra Section: Spaces of Differentiable Functions 57
That is, Cbm (R) is the space of m-times differentiable functions such that
f and each derivative f ′ , . . . , f (m) is continuous and bounded, and so forth.
According to Problem 3.50.53, these spaces are nested in the following way:
The inclusions on the preceding line are proper. For example, the function
f (x) = sin x belongs to Cbm (R) because it and every derivative is bounded,
but f ∈/ C0m (R). Problem 3.50.53 asks for examples that show that the other
inclusions in equation (3.34) are also proper.
After giving some examples, we will define the “standard norm” for the
subspaces Cbm (R), C0m (R), and Ccm (R), and determine which of these are
complete with respect to that norm. The hardest case is actually m = 1;
once we understand that case we can extend to higher orders by induction.
Sometimes it is useful to consider analogous spaces of infinitely differen-
tiable functions. Although these are not normed spaces, we make the following
definitions:
C ∞ (R) = f ∈ C(R) : f, f ′ , . . . ∈ C(R) ,
Cb∞ (R) = f ∈ Cb (R) : f, f ′ , . . . ∈ Cb (R) ,
C0∞ (R) = f ∈ C0 (R) : f, f ′ , . . . ∈ C0 (R) ,
Cc∞ (R) = f ∈ Cc (R) : f, f ′ , . . . ∈ Cc (R) .
It is not obvious that Cc∞ (R) 6= {0}. Some examples of nontrivial elements
of Cc∞ (R) are constructed in Problem 2.10.6.
There are many ways to define a norm on Cb1 (R). For example, since Cb1 (R)
is a subspace of Cb (R), the uniform norm is a norm on Cb1 (R). However,
the following exercise shows that Cb1 (R) is not complete with respect to the
uniform norm.
Exercise 3.50.51. The hat function or tent function on the interval [−1, 1]
is the piecewise linear function W pictured in Figure 3.57. Explicitly,
1 − x, 0 ≤ x ≤ 1,
W (x) = max 1 − |x|, 0 = 1 + x, −1 ≤ x ≤ 0, (3.35)
0, |x| ≥ 1.
W
0.5
1
f5
0.5
-0.5
-1
g5
0.5
Fig. 3.59 The function g5 from Example 3.50.51, with the hat function W in dashed lines
for comparison.
In summary, k · ku is a norm on Cb1 (R), but Cb1 (R) is not complete with
respect to this norm. We need to construct a different norm that incorporates
information about the derivative. As a first attempt, let us define
Unfortunately, this does not even define a norm on Cb1 (R). For example, the
constant function f = 1 satisfies
|||f ||| = kf ′ ku = 0,
even though f is not the zero function. Hence ||| · ||| does not satisfy the
uniqueness requirement of a norm.
The “correct” definition of a norm on Cb1 (R) should incorporate informa-
tion about both f and f ′ . We can do this by setting
According to Problem 3.5.25, this is a norm on Cb1 (R), and Cb1 (R) is com-
plete with respect to this norm.
60 Extra Material and Problems c 2018 Christopher Heil
The case m = 1 (which is Problem 3.5.25) is actually the hard part. As-
suming that problem has been done, we can extend to higher-order derivatives
by induction. Given m ∈ N, the norm on Cbm (R) is defined by
Exercise 3.50.52. Use induction to prove that Cbm (R) is a Banach space
with respect to the norm k · kCbm . ♦
Problems 3.50.56 and 3.50.57 consider whether the spaces C0m (R) and
Ccm (R) are complete with respect to the norm k · kCbm (yes for C0m (R), but
no for Ccm (R)).
Extra Problems
Prove that the function γ constructed in Problem 2.10.6 is not real analytic
at the origin.
3.50.55. Define
( 2
−1/x2 e−1/x , x > 0,
γ(x) = e χ(0,∞) (x) =
0, x ≤ 0.
(c) If a < b then the function f (x) = γ(x − a) γ(b − x) belongs to Cc∞ (R)
and satisfies f (x) > 0 for x ∈ (a, b) and f (x) = 0 for x ∈
/ (a, b).
3.50.56. Given m ∈ N, prove that C0m (R) is a Banach space with respect to
the norm k · kCbm .
3.50.57. Let g be any function in C0m (R) such that g ∈ / Ccm (R). Let θ be
the function constructed in Problem 2.10.7 (so θ is infinitely differentiable,
compactly supported, and identically 1 on [−1, 1]). For each k ∈ N, set
is finite for every f ∈ C(R). In fact, considering a single, fixed N, the function
||| · |||N is nonnegative, homogeneous, and satisfies the Triangle Inequality, so
it is a seminorm on C(R) in the sense of Definition 3.1.1. However, it is not
a norm, because there exist functions in C(R) other than the zero function
that satisfy |||f |||N = 0.
Remark 3.51.50. Let χ[−N,N ] denote the characteristic function of the inter-
val [−N, N ], i.e., (
1, |x| ≤ N,
χ[−N,N ] (x) =
0, |x| > N.
Then f · χ[−N,N ] is the function that equals f on [−N, N ] and is zero outside
of [−N, N ]. Therefore, an equivalent way to write ||| · |||N is
That is, the seminorm |||f |||N is simply the uniform norm of the restriction
of f to the interval [−N, N ]. If f is zero on [−N, N ] but nonzero somewhere
outside of that interval, then we will have |||f |||N = 0 even though f is not
the zero function. ♦
We would like to collect all of these seminorms ||| · |||N together to create a
norm or metric on C(R). We could try summing over N, but unfortunately
the series ∞ ∞
X X
|||f |||N or even 2−N |||f |||N
N =1 N =1
need not converge for every f ∈ C(R). On the other hand, no matter what
f ∈ C(R) we choose we always have
|||f |||N
≤ 1, all N ∈ N,
1 + |||f |||N
so we can define
∞
X |||f |||N
kf k = 2−N , f ∈ C(R).
1 + |||f |||N
N =1
3.51 Extra Section: A Metric for C(R) 63
This series converges for every f ∈ C(R), but does it define a norm? Unfortu-
nately, the answer is no, because kcf k 6= |c| kf k in general. Even so, Problem
3.5.70 shows that
∞
X |||f − g|||N
d(f, g) = kf − gk = 2−N , f, g ∈ C(R), (3.39)
1 + |||f − g|||N
N =1
is a metric on C(R).
What kind of convergence criterion does the metric d correspond to? That
is, what does it mean for fk to converge to f with respect to the metric d?
Instead of uniform convergence, we will see that convergence with respect to
d is the type of convergence that is given in the following definition.
gn
-n n n+1
On the other hand, fk does not converge uniformly to f, because for every
k we have
64 Extra Material and Problems c 2018 Christopher Heil
The following lemma shows that convergence with respect to the metric d
is precisely the same as uniform convergence on compact sets.
lim d(fk , f ) = 0.
k→∞
= 2N d(f − fk )
< 2N η
ε
= .
1+ε
x
Since the function 1+x is monotone increasing on (0, ∞), it follows that for
all k > K we have
|||f − fk |||N < ε.
This proves that |||f − fk |||N → 0 as k → ∞.
3.51 Extra Section: A Metric for C(R) 65
M
X ∞
X
ε
≤ 2−N + 2−N · 1
1+0
N =1 N =M+1
∞
X
≤ ε 2−N + ε
N =1
= 2ε.
Hence d(fk , f ) → 0 as k → ∞. ⊓
⊔
Example 3.51.52 shows that a sequence that converges uniformly on com-
pact sets need not converge uniformly. On the other hand, if fk converges
uniformly to f then we have kf − fk ku → 0 as n → ∞. Therefore, for any
particular N,
= kf − fk ku
→ 0 as k → ∞.
1.0
W5 W6
0.8
0.6
0.4
0.2
0.0
0 1 2 3 4 5 6 7
Fig. 3.61 The functions Wk , k ≥ 5, from Example 3.51.54, vanish on the interval [−4, 4].
The function
Wk (x) = W (x − k)
is simply the hat function translated to the right by k units. If we fix an
integer N > 0, then for every k > N we have Wk = 0 on the interval
[−N, N ]. Consequently, as shown in Figure 3.61, kWk · χ[−N,N ] ku = 0 for
k > N. Therefore
Extra Problems
Remark: Using the terminology of Definition 2.6.3, this says that Cc (R) is
dense in C(R) with respect to the metric d.
3.51.56. Prove that C(R) is complete with respect to the metric d defined
in equation (3.39).
Prove that
∞ X
X ∞
|||f − g|||N,n
d(f, g) = 2−n−N , f, g ∈ C ∞ (R),
n=0 N =1
1 + |||f − g|||N,n
We saw in Example 2.9.53 that a continuous function need not map an open
set to an open set. We are only guaranteed that the inverse image of an open
set under a continuous function is open. Now, if f : X → Y is a bijection, then
it has an inverse function f −1 : Y → X. If the inverse function is continuous,
then the inverse image of an open set under f −1 will be open. As we observe
in the following lemma, the inverse image of a set under f −1 is simply the
direct image of the set under f. So if f and f −1 are both continuous, then
both direct and inverse images of open sets will be open. Here is the precise
statement and proof of this fact.
Lemma 3.52.50. Let X and Y be metric spaces, and suppose that f : X → Y
is a bijection. If f and its inverse function f −1 are both continuous, then
and
V is open in Y =⇒ f −1 (V ) is open in X. (3.41)
Proof. Equation (3.41) is an immediate consequence of the fact that f is
continuous. To prove equation (3.40), let U be any open subset of X. Then,
since f −1 : Y → X is continuous, the inverse image of U under the function
f −1 must be open. Because f is a bijection, this inverse image is simply
(f −1 )−1 (U ) = U.
Therefore U is open in X. ⊓
⊔
Thus, if f and f −1 are both continuous, then f and f −1 both preserve
topological structure, as they each map open sets to open sets. We have a
special name for such functions.
Definition 3.52.51 (Homeomorphism). Let X, Y be metric spaces.
(a) We say that f : X → Y is a homeomorphism if f is a bijection and both
f and its inverse function f −1 are continuous.
(b) We say that X and Y are homeomorphic metric spaces if there exists a
homeomorphism f : X → Y. ♦
If f : X → Y is a homeomorphism, then f maps the open sets in X to the
open sets in Y, and f −1 does the same in the reverse direction. In essence,
homeomorphic spaces have the “same topological structure.” The following
exercise gives an example.
Exercise 3.52.52. Define f : R2d → Cd by
Example 3.52.53. For each fixed real number a ∈ R we will define a function
Ta : Cb (R) → Cb (R). That is, Ta will maps elements of Cb (R) to elements
of Cb (R). In other words, Ta is a kind of “meta-function” that inputs a
function g ∈ Cb (R) and outputs a new function in Cb (R). Using normal
function notation we would denote this output by Ta (g), but in order to
avoid multiplicities of parentheses we will simply write Ta g instead of Ta (g).
We need to declare the rule for Ta , i.e., we must define how Ta will take
the input g and determine what the corresponding output Ta g is. What we
have in mind for Ta is that it will simply translate the graph of the input g
by a units to give us Ta g. That is, given g ∈ Cb (R), we declare that Ta g is
the function whose graph is the identical to the graph of g except that it has
been shifted to the right by a units. Explicitly,
1.0
0.8
g 0.6 T5 g
0.4
0.2
-4 -2 2 4 6 8 10
Ta f (t) = f (t − a) = f (x)
and similarly
Ta g(t) = g(t − a) = g(x).
But Ta f (t) = Ta g(t), so this implies that f (x) = g(x). Since x was an ar-
bitrary real number, we have shown that f and g take the same values at
every x, and therefore they are they same function. That is, f = g. Therefore
we have proved that Ta is injective.
Proof that Ta is surjective. We must prove that the range of Ta is all
of Cb (R). So, choose any function g ∈ Cb (R). We must show that f is in
the range of Ta . In other words, we must prove that g = Ta f for some
function f, i.e., we want g to be the function f shifted a units to the right.
The appropriate function f is therefore g shifted a units to the left. That is,
f is the function f (x) = g(x + a). This function f belongs to Cb (R), and for
every x we have
Ta f (x) = f (x − a) = g(x).
Note that this discussion also reveals what (Ta )−1 is. The inverse of shifting a
units to the right is simply shifting a units to the left, so
kTa gku = sup |Ta g(x)| = sup |g(x − a)| = sup |g(t)| = kgku. (3.42)
x∈R x∈R t∈R
lim kg − gn ku = 0.
n→∞
But then, by the translation-invariance of the uniform norm and the fact that
Ta (g − gn ) = Ta g − Ta gn , we see that
3.52 Extra Section: Homeomorphisms 71
= lim kg − gn ku
n→∞
= 0.
In Section 6.1 we will see much simpler ways to prove that a linear function
is continuous.
Remark 3.52.54. We will show in Theorem 6.3.1 that if X and Y are normed
spaces and X is finite-dimensional, then every linear function F : X → Y is
continuous. However, if X is infinite-dimensional, then there exist linear func-
tions F : X → Y that are not continuous! An example of such a discontinuous
linear function appears in Example 6.3.4. ♦
Extra Problems
Prove that L : Cb (R) → Cb (R) is continuous and injective, but is not a home-
omorphism.
(b) Now define M : Cb (R) → Cb (R) by
72 Extra Material and Problems c 2018 Christopher Heil
Fz (x) = x + z, x ∈ X.
Da f (x) = f (ax), x ∈ R.
Since
2N − 1
x − sN = x − x = 2−N x,
2N
we have
XN
(−1)n
tN =
n=1
n
XN XN
(−1) (−1)n
sN = δ1 = , 0, 0, . . . = (tN , 0, 0, . . . ).
n=1
n n=1
n
kx − sN k1 = (− ln 2 − tN , 0, 0, . . . ) 1
= | − ln 2 − tN | → 0 as N → ∞.
74 Extra Material and Problems c 2018 Christopher Heil
P∞ n
Hence sN does converge to x in ℓ1 -norm, so n=1 (−1)n δ1 converges in ℓ
1
p
and equals x. The reader should check that this series converges in ℓ for
every index 1 ≤ p ≤ ∞. ♦
Extra Problems
Let {xn }n∈N be a sequence of vectors in a normed space X. Then {kxn k}n∈N
is a sequence of real scalars, and we canPwonder whether there is a connection
between the convergencePof the series xn in X and the convergence of the
series of real numbers kxn k. To address this, let us give names to the
partial sums of these two series, say
N
X N
X
sN = xn and tN = kxn k.
n=1 n=1
It follows from the Triangle Inequality that ksN k ≤ |tN |, but this inequality
does not give us any useful information about convergence. On the other
hand, if M < N then we can use the Triangle Inequality to compute that
N
X N
X
ksN − sM k = xn ≤ kxn k = |tN − tM |.
n=M+1 n=M+1
∞
X
kxn k converges =⇒ {tN }N ∈N is Cauchy in R (4.44)
n=1
=⇒ {sN }N ∈N is Cauchy in X (4.45)
∞
X
=⇒ xn converges. (4.46)
if X is complete
n=1
c1 = 3,
c2 = 0.1,
c3 = 0.04,
c4 = 0.001,
c5 = 0.0005,
P P
and so forth. Since |cn | < ∞, the series cn converges absolutely. How-
P
ever, the series cn does not converge in the set Q. The partial sums
PN
sN = n=1 c n are rational numbers that converge to the real number π,
but the limit π does not belong to Q. ♦
Extra Problems
4.2.51. Give an explicit example of functions fn ∈ Cc (R) such that the series
P ∞
n=1 fn converges absolutely in Cc (R) with respect to the uniform norm,
but the series does not converge to an element of Cc (R).
76 Extra Material and Problems c 2018 Christopher Heil
|f (x) − f (y)|
|||f ||| = |f (x0 )| + sup
x6=y |x − y|
does converge (in fact, its partial sums converge to ln 12 , the natural loga-
rithm of 12 ). Although the alternating harmonic series converges, it does not
converge absolutely in the sense of Definition 4.2.2.
We will show that the alternating harmonic series does not converge un-
1 1
conditionally either. To do this, set pn = 2n−1 and qn = 2n , i.e., the pn are
the positive terms from the alternating series and the qn are the absolute
P P
values of the negative terms. Each series pn and qn diverges to infinity.
4.3 Unconditional Convergence of Series 77
p1 + · · · + pm1 > 1.
X∞
(−1)σ(n)
n=1
σ(n)
converges for some bijections σ : N → N but not for others, which says that
the alternating harmonic series converges conditionally. ♦
Extra Problems
P
4.3.56. Let xn = (−1)n/n, so xn is the alternating harmonic series. Al-
though
P this series converges, show that there exist scalars cn such that cn → 0
yet cn xn does not converge.
P
4.3.57. Assume that cn is a conditionally convergent series of real scalars,
i.e., the series converges, but it does not converge unconditionally.
(a) Let (pn ) be the sequence of nonnegative terms of (cn ) in order,P and let
(q
P n ) be the sequence of negative terms of
P (c n ) in order. Show that pn and
qn must both diverge. Conclude that cn is not absolutely convergent.
P
(b) Given x ∈ R, show there exists a permutation σ of N such that cσ(n)
converges and equals x.
P
(c) Show that there exists a permutation σ of N such that cσ(n) diverges
PN
to ∞ (that is, limN →∞ n=1 cσ(n) = ∞), and another permutation τ such
P
that cτ (n) diverges to −∞.
P
(d) Show that there exists a permutation σ of N such that cσ(n) does
not converge and does not diverge to ∞ or −∞.
4.5 Independence, Hamel Bases, and Schauder Bases 81
P
4.3.58. Let X be a finite-dimensional normed space. Show that a series xn
in X is unconditionally convergent if and only if it is absolutely convergent.
P
4.3.59. Use Theorem 4.3.55 to show that if a series ∞ n=1 xn converges un-
conditionally in a Banach space X, then there P exists a single vector x ∈ X
∞
such that for each bijection σ : N → N we have n=1 xσ(n) = x.
4.3.60. Let XPbe P a Banach space, and assume that vectors xmn ∈ X, m,
n ∈ N, satisfy m n kxmn k < ∞. Show that if σ : N → N×N is a bijection,
then the following series all converge and are equal as indicated:
∞ X
X ∞ ∞ X
X ∞ ∞
X
xmn = xmn = xσ(k) .
m=1 n=1 n=1 m=1 k=1
4.3.61. Prove that the conclusion of Problem 4.3.60 remains valid if we re-
place the hypothesis of absolute
Pconvergence with unconditional convergence,
i.e., we assume that the series (m,n)∈N2 cmn converges unconditionally in X.
When we are given a finite-dimensional vector space, one of the first things
that we are taught to do in undergraduate linear algebra class is to look for
a basis for that space, i.e., a set that both spans and is linearly independent.
Unfortunately, bases in this ordinary vector space sense are usually not very
useful in the infinite-dimensional setting. For one thing, in most infinite-
dimensional vector spaces the only way that we can know that a basis exists
is by appealing to the Axiom of Choice. For another, any basis for a complete
infinite-dimensional normed space must be an uncountable set (see Theorem
4.5.3). This makes them too unwieldy for practical use.
When we work with generic vector spaces, we have no choice but to restrict
ourselves to finite linear combinations of elements. In particular, spanning
and linear independence are each defined in terms of finite linear combinations
of vectors. We cannot form an infinite series in a vector space without having
a notion of what it means to converge in that space, because an infinite series
is, by definition, the limit of the partial sums of the series. If our vector space
is lucky enough to have a norm, then we do have a convergence criterion and
therefore can form infinite series. Consequently, when we work with a normed
vector space we can seek “bases” that are defined in terms of “infinite linear
combinations” instead of just finite linear combinations of vectors. This is not
an issue in finite-dimensional normed spaces, but it can be very important
when dealing with infinite-dimensional spaces. As it turns out, defining the
most useful “basis” concept in infinite dimensions is not as straightfoward as
we might hope, as there are more subtleties in the ways that we can generalize
82 Extra Material and Problems c 2018 Christopher Heil
The standard basis vectors, which were introduced in Notation 2.1.8, will
be used to illustrate many of the concepts discussed in this chapter, so we
recall their definition here. Given n ∈ N, the nth standard basis vector is the
sequence δn whose nth component is 1, and all other components are zero.
That is,
δn = (0, . . . , 0, 1, 0, 0, . . . ),
where n − 1 zeros precede the 1. The sequence of standard basis vectors, or
simply the standard basis, is the collection E of all of the standard basis
vectors. In other words,
E = {δn }n∈N .
Each standard basis vector δn has only finite many nonzero components, so
δn belongs to the space c00 . Hence the standard basis E is a subset of c00 ,
and since c00 is contained in the spaces ℓp and c0 , the standard basis E is also
a subset of ℓp and c0 .
Although we have just called E = {δn }n∈N a “basis,” this should only
be viewed as a name for now. As we progress through the chapter, we will
discuss in what sense E is a basis for spaces such as ℓp , c0 , or c00 .
According to the next exercise, if p is finite then the standard basis is
a Schauder basis for ℓp . Note that in order to prove that a sequence is a
Schauder basis for a given space, we must show that the series in equation
(4.17) converge in the norm of that space (not just in some other sense like
componentwise convergence!).
Closed Span
We emphasize that the definition of the closed span does not say that
4.5 Independence, Hamel Bases, and Schauder Bases 83
X
∞
span(A) = cn xn : xn ∈ A, cn ∈ F ← This need not hold!
n=1
n N
X o
span{xn }n∈N = x ∈ X : ∃ cn,N ∈ F such that cn,N xn → x as N → ∞ .
n=1
That is, x belongs to the closed span of {xn }n∈N if and only if there exist
scalars cn,N ∈ F such that
N
X
cn,N xn → x as N → ∞. (4.47)
n=1
The scalars cn,N in equation (4.47) can depend on N. In contrast, to say that
P∞
x = n=1 cn xn means that
N
X
cn xn → x as N → ∞. (4.48)
n=1
Schauder Bases
We will prove that every Schauder basis is a complete sequence and is finitely
linearly independent. However, Example 4.5.52 will show that the converse
statement need not hold.
N
X
cn xn = 0
n=1
Example 4.5.52. The set of monomials M = {xk }∞ k=0 is both complete and
finitely independent in C[a, b]. However, as shown in Problem 4.6.6,Pthere ex-k
ist functions f ∈ C[a, b] that cannot be written in the form f (x) = ∞k=0 ck x
with convergence of the series in the uniform norm (or even just pointwise).
Consequently the set of monomials M is not a Schauder basis for C[a, b]. ♦
Exercise 4.5.53. Let X be a Banach space. Show that if there exists a count-
able sequence {xn }n∈N that is complete in X, then
X
N
S = rn xn : N > 0, rn is rational
n=1
For example, we can use this exercise to prove that C[a, b] is separable.
Thus, a Banach space that has a Schauder basis must be separable. The
question of whether every separable Banach space has a Schauder basis was
a longstanding open problem known as the Basis Problem. It was finally
shown by Enflo [Enf73] that there exist separable Banach spaces that have no
Schauder bases! For constructions of Schauder bases in some specific Banach
spaces such as C[0, 1], we refer to the text [Heil11].
The problem given in Problem 4.5.12 is one particular open problem that is
part of the Linear Independence of Time-Frequency Translates Conjecture,
also known as the HRT Conjecture. Many special cases are known. For ex-
ample, if g ∈ Cc (R) then it is easy to prove that the set of four functions
given in equation (4.21) is independent. More special cases can be found in
the survey paper [Heil06]. The original conjecture is in the paper:
C. Heil, J. Ramanathan, and P. Topiwala, Linear independence of
time-frequency translates, Proc. Amer. Math. Soc., 124 (1996), pp. 2787–
2795.
The actual statement of the conjecture is as follows.
86 Extra Material and Problems c 2018 Christopher Heil
2
Conjecture 4.5.56 (HRT Conjecture). If g in L (R) is not the zero function
and Λ = (ak , bk ) : k = 1, . . . , N is a set of finitely many distinct points in
R2 , then
G(g, Λ) = e2πibk x g(x − ak ) : k = 1, . . . , N
is linearly independent. ♦
Extra Problems
Since
c00 ( c0 ( c ( ℓ∞ ,
the sup-norm is a norm of each of these four spaces. Further, by Problem
2.2.19, ℓ∞ is complete with respect to this norm. Prove the following state-
ments.
(a) Prove that c is a closed subspace of ℓ∞ , and therefore is a Banach
space with respect to the sup-norm.
(b) c0 is a closed subspace of c, and therefore is a Banach space with
respect to the sup-norm.
(c) c00 is not a closed subspace of c0 , and therefore is not complete with
respect to the sup-norm.
is dense in C0 (R).
Extra Problems
There are many statements that are equivalent to the Axiom of Choice.
In the remainder of this appendix, we will discuss another equivalent form
of the Axiom of Choice, known as Zorn’s Lemma or the Kuratowski–Zorn
Lemma. First we need to introduce the concept of a partial order on a set.
B is comparable to A =⇒ B ≤ A.
Proof. Let
S = A ⊆ X : A is finitely linearly independent .
is also finitely independent. To see this, choose finitely many distinct vectors
90 Extra Material and Problems c 2018 Christopher Heil
Extra Problems
5.7.50. This problem will give a constructive solution to Exercises 3.7.51 and
3.7.52 in the setting of Hilbert spaces.
(a) Let M be a proper, closed subspace of a Hilbert space H. Given a vector
f∈/ M, let p be the orthogonal projection of f onto M and set e = f − p.
Show that the vector g = e/kek satisfies
and therefore {en } is a Schauder basis for H. This fact is very useful. Not
every complete sequence {fn } is a Schauder basis, and even if it is, it can
be a highly nontrivial
P task to find an explicit formula for the scalars cn (f )
that satisfy f = cn fn . Remarkably, if {en } is both complete and ortho-
normal, then it is a automatically a Schauder basis with explicitly known
representations. Furthermore, such an orthonormal basis has a number of
important “stability”
P properties. First, it follows from Theorem 5.7.1 that
the series f = hf, en i en converges unconditionally. That is, the series con-
verges no matter how we order the index set (and other important “stability”
features of unconditional convergence appear in Theorem 4.3.55). Second, the
92 Extra Material and Problems c 2018 Christopher Heil
Extra Problems
this chapter on orthonormal sequences carry over with only minor notational
changes to sequences indexed by Z or any other countably infinite index set.
In particular, Theorems 5.7.1 and 5.8.1 both carry over to orthonormal se-
quences that are indexed by any countable index set. Essentially, this is a
consequence of the unconditional convergence of the series that appear in
those theorems. An unconditionally convergent series converges regardless
of what ordering we impose on the index set, and hence we can use any
countably infinite index set that we like without changing the meaning of the
series.
For the index set Z, we typically use the “standard ordering”
0, 1, −1, 2, −2, 3 − 3 . . . .
s1 = x0 ,
s2 = x0 + x1 ,
s3 = x0 + x1 + x−1 ,
s4 = x0 + x1 + x−1 + x−2 ,
P
and so forth. Therefore, using this ordering, an infinite series n∈Z xn con-
verges if and only if these partial sums sN converge as N → ∞. If the series
converges unconditionally (the typicaly situation when working with ortho-
normal vector), then it will converge no matter what ordering of the series
that we choose. When dealing with bi-infinite series it is often convenient
to consider the partial sum s2N −1 = x−N + · · · + xN . We call this the N th
symmetric partial sum and often denote it by
N
X
SN = s2N −1 = xn = x0 + x1 + x−1 + · · · + xN + x−N .
n=−N
Extra Problems
5.11.50. Let f (x) = x for x ∈ [0, 1]. Compute the Fourier coefficients of f,
and use this to give another proof of Euler’s formula.
6.4 Equivalence of Bounded and Continuous Linear Operators 95
Functionals
If the codomain of an operator is the field of scalars F, then we say that the
operator is a functional. In other words, a functional is simply a function
of the form f : X → F. We often denote functionals by Greek letters such
as λ, µ, or ρ, or sometimes by lowercase Roman letters such as d or q. The
following exercise gives two examples of nonlinear functionals. We will see
some linear functionals in Example 6.3.2.
Exercise 6.1.50. (a) Let X be a nontrivial normed space. Define a func-
tional ρ : X → F by
ρ(x) = kxk, x ∈ X.
Prove that ρ is not linear, not surjective, and not injective, even though
ker(ρ) = {0}.
(b) Fix n ≥ 2 and let X be the set of all n × n matrices with scalar entries.
Let det(A) denote the determinant of a matrix A, and define d : X → F
by d(A) = det(A). Show that the functional d is surjective, but it is not
linear and not injective. ♦
Recall from Exercise 2.9.52 that if X and Y are normed spaces, then a func-
tion A : X → Y is continuous at a point x ∈ X if xn → x in X implies
Axn → Ax in Y, and A is continuous if it is continuous at every point.
Here is an expanded version of Theorem 6.4.1.
(b) A is continuous at x = 0.
(c) A is continuous.
(d) A is bounded.
Proof. (a) ⇒ (b), (c). Suppose that A is continuous at x, and fix any vector
y ∈ X. If yn → y then y − yn + x → x, so by linearity and the definition of
continuity at f we have
Ay − Ayn + Ax = A(y − yn + x) → Ax as n → ∞.
kxn k
kyn − 0k = kyn k = → 0 as n → ∞.
n
Using abstract language, Lemma 6.5.5 shows that B(X) is a Banach algebra.
This Banach algebra is noncommutative since AB 6= BA in general, but on
the other hand it does have an identity element (the identity operator I).
6.5 The Space B(X, Y ) 97
Extra Problems
6.5.51. (a) What is the difference between Cb (R) and B(R, F)? Is one a subset
of the other?
(b) Give a complete characterization of the elements of B(R, F).
(b) Prove that if the norm on Fn and Fm is k · k∞ then the operator norm
of A is X
n
kAkℓ →ℓ = max
∞ ∞ |aij | .
i=1,...,m
j=1
Extra Problems
6.6.50. (a) Prove that the dilation operator Da : Cb (R) → Cb (R) defined
in Problem 3.52.58 is a bounded linear bijection, and its operator norm is
kDa k = 1. Is it an isometry?
(b) Prove that the operators L and M defined in Problem 3.52.56 are
bounded and linear, and find their operator norms. Is either one an isometry?
6.6.51. (a) We say that a function f : R → F is 1-periodic if f (x + 1) = f (x)
for all x ∈ R. Prove that
Cbper (R) = f ∈ Cb (R) : f is 1-periodic
is a closed subspace of C[0, 1] with respect to the uniform norm on C[0, 1].
(c) Prove that
Cbper (R) ∼
= C per [0, 1],
i.e., there exists an isometric isomorphism A : Cbper (R) → C per [0, 1].
6.6.52. Suppose that X and Y are normed spaces, and {xn }n∈N is a Schauder
basis for X. Show that if A : X → Y is a topological isomorphism, then
{Axn }n∈N is a Schauder basis for Y.
If the series in equation (6.52) does not converge for some k, then the convo-
lution of x with y is undefined. ♦
The convolution of two arbitrary sequences need not be defined. For ex-
ample, if
x = 2k k∈Z and y = 2−k k∈Z ,
then the series given in equation (6.52) does not converge for any k, so the
convolution of these two particular sequences does not exist. If the convolu-
tion x ∗ y does exist, then we usually denote the kth component of x ∗ y by
(x ∗ y)k instead of ck . That is, if x ∗ y exists then its components are
∞
X
(x ∗ y)k = xj yk−j , k ∈ Z.
j=−∞
The following theorem shows that the convolution of two summable se-
quences always exists, and it also gives us a bound for the ℓ1 -norm of a
convolution. In the proof, we use the fact that we can reindex convergent
series. In particular, if k is fixed then by setting i = k − j we see that
∞
X ∞
X
yk−j = yi .
j=−∞ i=−∞
Note that it is important here that the summation is over all of Z, otherwise
the range of summation would change when we reindex. This is one reason
why we use bi-infinite sequences when we define convolution. The proof of
the following result is Problem 6.7.9.
Thus ℓ1 (Z) is closed under convolution. Problem 6.7.10 gives some further
properties of convolution.
In summary, not only is ℓ1 (Z) a Banach space, but it is also closed under
the operation of convolution. Indeed, convolution is a “multiplication-like”
operation on ℓ1 (Z) in the sense that it has properties similar to ordinary
multiplication of scalars. We have a special name for normed spaces that
have such a “multiplicative” operation.
6.7 Infinite Matrices 101
Some normed algebras also have an additional operation that has proper-
ties similar to that of conjugation of complex numbers.
Extra Problems
6.7.56. (a) Let c00 (Z) be the set of all bi-infinite sequences x = (xk )k∈Z such
that xk 6= 0 for at most finitely many k. Prove that c00 (Z) is closed under
convolution, i.e., if x, y ∈ c00 (Z) then x ∗ y exists and belongs to c00 .
(b) Let c0 (Z) be the set of all bi-infinite sequences x = (xk )k∈Z such that
xk → 0 as k → ±∞. Exhibit sequences x, y ∈ c0 (Z) such that the convolution
x ∗ y does not exist.
6.7.58. Fix 1 < p < ∞, and suppose that x ∈ ℓp (Z) and y ∈ ℓ1 (Z).
(a) Show that
∞
X
1/p 1/p′
|(x ∗ y)k | ≤ |xj | yk−j yk−j dy. (6.54)
j=−∞
Apply Hölder’s Inequality with exponents p and p′ to the two factors that
appear on the right-hand side of equation (6.54) to show that
X
∞ 1/p
1/p′ p
|(x ∗ y)k | ≤ kyk1 |xj | |yk−j | .
j=−∞
(b) Use part (a) and Tonelli’s Theorem to directly prove Young’s Inequal-
ity:
kx ∗ ykp ≤ kxkp kgk1 . (6.55)
Remark: Equation (6.55) also holds for p = 1 (see Problem 6.7.9) and for
p = ∞ (see Problem 6.7.11).
7.8 The Spectral Theorem for Compact Self-Adjoint Operators 103
Extra Problems