a brief guide to metrics, norms, and inner...

70
Christopher Heil A Brief Guide to Metrics, Norms, and Inner Products An Electronic Manuscript April 12, 2019 c 2016 Christopher Heil

Upload: others

Post on 23-Mar-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Christopher Heil

A Brief Guide to Metrics, Norms,

and Inner Products

An Electronic Manuscript

April 12, 2019

c©2016 Christopher Heil

Page 2: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric
Page 3: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Contents

1 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Metrics and Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Convergence and Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.3 Topology in Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.4 Compact Sets in Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 91.5 Continuity for Functions on Metric Spaces . . . . . . . . . . . . . . . . . 131.6 Urysohn’s Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 Norms and Banach Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.1 The Definition of a Norm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.2 The Induced Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.3 Properties of Norms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.4 Convexity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.5 Banach Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.6 Infinite Series in Normed Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 242.7 Span and Closed Span . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.8 Hamel Bases and Schauder Bases . . . . . . . . . . . . . . . . . . . . . . . . . 272.9 The Space Cb(X) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3 Inner Products and Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . . . 373.1 The Definition of an Inner Product . . . . . . . . . . . . . . . . . . . . . . . 373.2 Properties of an Inner Product . . . . . . . . . . . . . . . . . . . . . . . . . . . 383.3 Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393.4 Orthogonal and Orthonormal Sets . . . . . . . . . . . . . . . . . . . . . . . . 443.5 Orthogonal Complements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443.6 Orthogonal Projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453.7 Orthonormal Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483.8 Orthonormal Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523.9 Existence of an Orthonormal Basis . . . . . . . . . . . . . . . . . . . . . . . 54

Index of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

v

Page 4: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

vi Contents

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

Page 5: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Chapter 1

Metric Spaces

We will introduce and quickly study the basics of metrics, norms, and in-ner products in this short manuscript. This manuscript should be accessibleto readers who have a background in undergraduate real analysis topics, in-cluding sequences, limits, suprema and infima, functions, cardinality, infiniteseries, Riemann integration, and vector spaces. Assuming that background,the manuscript essentially self-contained, although the presentation is quitecompressed. Most major proofs are included, while certain other proofs areassigned as problems, and references are provided for proofs that are omitted.Explicit proofs of most of the statements made in this appendix, with exten-sive discussion and motivation, appear in the forthcoming volume [Heil16],but many of these can also be found in the volume [Heil11], which has alreadyappeared.

We assume that the reader is familiar with vector spaces (which are alsocalled linear spaces). The scalar field associated with the vector spaces inthis manuscript will always be either the real line R or the complex plane C.When we specifically deal with the vector space Rd, we assume that the scalarfield is R, and likewise we always assume that the scalar field for Cd is C.For simplicity of presentation, when dealing with a vector space we oftenimplicitly assume that the scalar field is C, but usually only minor (andobvious) changes are required if the scalar field is R. The elements of thescalar field are often called scalars, so for us a scalar will always mean a realor complex number.

1.1 Metrics and Convergence

A metric provides us with a notion of the distance between points in a set.

Definition 1.1.1 (Metric Space). Let X be a nonempty set. A metric onX is a function d: X × X → R such that for all x, y, z ∈ X we have:

c©2014 Christopher Heil1

Page 6: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2 1 Metric Spaces

(a) Nonnegativity: 0 ≤ d(x, y) < ∞,

(b) Uniqueness: d(x, y) = 0 if and only if x = y,

(c) Symmetry: d(x, y) = d(y, x), and

(d) The Triangle Inequality: d(x, z) ≤ d(x, y) + d(y, z).

If these conditions are satisfied, then X is a called a metric space. The numberd(x, y) is called the distance from x to y. ♦

A metric space X does not have to be a vector space, although most ofthe metric spaces that we will encounter in this manuscript will be vectorspaces (indeed, most are actually normed spaces). If X is a generic metricspace, then we often refer to the elements of X as “points,” but if we knowthat X is a vector space, then we may refer to the elements of X as “vectors.”We will mostly use letters such as x, y, z to denote elements of a metric ornormed space. However, in many cases we know that our space is a space offunctions. In such a context we usually use letters such as f, g, h to denoteelements of the space, and we may refer to those elements as “functions,”“vectors,” or “points.”

Here is an example of a metric on an infinite-dimensional vector space.

Example 1.1.2. Given a sequence of real or complex scalars

x = (xk)k∈N = (x1, x2, . . . ),

we define the ℓ1-norm of x to be

‖x‖1 = ‖(xk)k∈N‖1 =∞∑

k=1

|xk|. (1.1)

We say that the sequence x is absolutely summable (or just summable forshort) if ‖x‖1 < ∞, and we let ℓ1 denote the space of all absolutely summablesequences, i.e.,

ℓ1 =

{

x = (xk)k∈N : ‖x‖1 =

∞∑

k=1

|xk| < ∞

}

.

This set is a vector space; in particular, it is closed under the operations ofaddition of sequences and multiplication of a sequence by a scalar. Further,a straightforward calculation shows that the function d defined by

d(x, y) = ‖x − y‖1 =

∞∑

k=1

|xk − yk|, x, y ∈ ℓ1, (1.2)

is a metric on ℓ1, so ℓ1 is both a vector space and a metric space. ♦

Page 7: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.2 Convergence and Completeness 3

The following particular elements of ℓ1 appear so often that we introducea name for them.

Notation 1.1.3 (The Standard Basis Vectors). Given an integer n ∈ N,we let δn denote the sequence

δn = (δnk)k∈N = (0, . . . , 0, 1, 0, 0, . . . ).

That is, the nth component of the sequence δn is 1, while all other componentsare zero. We call δn the nth standard basis vector, and we refer to the family{δn}n∈N as the sequence of standard basis vectors, or simply the standardbasis. ♦

1.2 Convergence and Completeness

Since a metric space has a notion of distance, we can define a correspondingnotion of convergence in the space, as follows.

Definition 1.2.1 (Convergent Sequence). Let X be a metric space. Asequence of points {xn}n∈N in X converges to the point x ∈ X if

limn→∞

d(xn, x) = 0.

That is, for every ε > 0 there must exist some integer N > 0 such that

n ≥ N =⇒ d(xn, x) < ε.

In this case, we write xn → x. ♦

Convergence implicitly depends on the choice of metric for X, so if we wantto emphasize that we are using a particular metric, we may write xn → xwith respect to the metric d.

Closely related to convergence is the idea of a Cauchy sequence, which isdefined as follows.

Definition 1.2.2 (Cauchy Sequence). Let X be a metric space. A se-quence of points {xn}n∈N in X is a Cauchy sequence if for every ε > 0 there

exists an integer N > 0 such that

m,n ≥ N =⇒ d(xm, xn) < ε. ♦

By applying the Triangle Inequality, we immediately obtain the followingrelation between convergent and Cauchy sequences.

Lemma 1.2.3 (Convergent Implies Cauchy). If {xn}n∈N is a convergentsequence in a metric space X, then {xn}n∈N is a Cauchy sequence in X.

Page 8: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

4 1 Metric Spaces

Proof. Assume that xn → x as n → ∞. Given ε > 0, there exists an N > 0such that d(x, xn) < ε

2 for all n > N. Consequently, if m, n > N then theTriangle Inequality implies that

d(xm, xn) ≤ d(xm, x) + d(x, xn) < ε. ♦

However, the converse of Lemma 1.2.3 does not hold in general.

Example 1.2.4. The “standard metric” on the real line R or the complex planeC is

d(x, y) = |x − y|.

Using the terminology we will make precise in Definition 1.2.7, each of R andC is a complete metric space with respect to this metric. This means thatany Cauchy sequence of real scalars must converge to a real scalar, and anyCauchy sequence of complex scalars must converge to a complex scalar (fora proof, see [Rud76, Thm. 3.11]).

Now consider the set of rational numbers Q. This is also a metric spacewith respect to the metric d(x, y) = |x − y|. However, we will show that Q

is not complete, i.e., we claim that there exist Cauchy sequences of rationalnumbers that do not converge to a rational number. For example, let xn bethe following rational numbers based on the decimal expansion of π:

x1 = 3, x2 = 3.1, x3 = 3.14, . . . .

Then {xn}n∈N is a Cauchy sequence in both Q and R. Moreover, it is aconvergent sequence in R because xn → π ∈ R. However, {xn}n∈N is not aconvergent sequence in the space Q, because there is no rational point r ∈ Q

such that xn → r. ♦

The metric space Q in Example 1.2.4 is not a vector space over the realfield. Here is an example of an infinite-dimensional vector space that containsCauchy sequences that do not converge in the space.

Example 1.2.5. Let c00 be the space of all sequences of real or complex num-bers that have only finitely many nonzero components:

c00 ={

x = (x1, . . . , xN , 0, 0, . . . ) : N > 0, x1, . . . , xN ∈ C

}

.

Since c00 is a subset of ℓ1, it is a metric space with respect to the metric ddefined in equation (1.2).

For each n ∈ N, let xn be the sequence

xn = (2−1, . . . , 2−n, 0, 0, 0, . . . ).

and consider the sequence of vectors {xn}n∈N, which is contained in both c00

and ℓ1. If m < n, then

Page 9: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.2 Convergence and Completeness 5

‖xn − xm‖1 =

n∑

k=m+1

2−k < 2−m,

and it follows from this that {xn}n∈N is a Cauchy sequence. If we considerthis sequence to be a sequence in the space ℓ1, then it does converge; in factxn → x where

x = (2−1, 2−2, . . . ) = (2−k)k∈N.

However, this vector x does not belong to c00, and there is no sequence y ∈ c00

such that xn → y. Therefore c00 is not a complete metric space. ⊓⊔

The set of rationals, Q, is not a vector space over the real field or thecomplex field (which are the only fields we are considering in this manuscript).In contrast, c00 is a vector space with respect to the real field (if we are usingreal scalars), or with respect to the complex field (if we are using complexscalars). In fact, c00 is the finite linear span of the set of standard basis vectorsE = {δk}k∈N:

c00 = span(E) =

{ N∑

k=1

xkδk : N > 0, x1, . . . , xN ∈ C

}

.

However, at least when we use the metric d defined by equation (1.2), thereare Cauchy sequences in c00 that do not converge to an element of c00. Thisfact should be compared to the next theorem, which shows that every Cauchysequence in ℓ1 converges to an element of ℓ1.

Theorem 1.2.6. If {xn}n∈N is a Cauchy sequence in ℓ1, then there exists avector x ∈ ℓ1 such that xn → x.

Proof. Assume that {xn}n∈N is a Cauchy sequence in ℓ1. Each xn is a vectorin ℓ1, and for this proof we will write the components of xn as

xn =(

xn(1), xn(2), . . .)

=(

xn(k))

k∈N.

By the definition of a Cauchy sequence, if we fix ε > 0, then there is aninteger N > 0 such that d(xm, xn) = ‖xm − xn‖1 < ε for all m,n > N.Therefore, if we fix a particular index k ∈ N then for all m, n > N we have

|xm(k) − xn(k)| ≤∞∑

j=1

|xm(j) − xn(j)| = ‖xm − xn‖1 < ε.

Thus, with k fixed,(

xn(k))

n∈Nis a Cauchy sequence of scalars and therefore

must converge. Definex(k) = lim

n→∞xn(k), (1.3)

and set x =(

x(1), x(2), . . .)

. For each fixed k, we have by construction thatthe kth component of xn converges to the kth component of x as n → ∞.

Page 10: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

6 1 Metric Spaces

We therefore say that xn converges componentwise to x. However, this is notenough. We need to show that x ∈ ℓ1, and that xn converges to x in ℓ1-norm.

To do this, again fix an ε > 0. Since {xn}n∈N is Cauchy, there is an N > 0such that ‖xm − xn‖1 < ε for all m, n > N. Choose any particular n > N,and fix an integer M > 0. Then, since M is finite,

M∑

k=1

|x(k) − xn(k)| =

M∑

k=1

limm→∞

|xm(k) − xn(k)|

= limm→∞

M∑

k=1

|xm(k) − xn(k)|

≤ limm→∞

‖xm − xn‖1

≤ ε.

As this is true for every M, we conclude that

d(xn, x) = ‖x − xn‖1 =∞∑

k=1

|x(k) − xn(k)|

= limM→∞

M∑

k=1

|x(k) − xn(k)|

≤ ε. (1.4)

Even though we do not know yet that x ∈ ℓ1, this tells us that the vec-tor x − xn has finite ℓ1-norm and therefore belongs to ℓ1. Since ℓ1 is closedunder addition, it follows that x = (x− xn) + xn ∈ ℓ1. Therefore our “candi-date sequence” x does belong to ℓ1. Further, equation (1.4) establishes thatd(xn, x) ≤ ε for all n > N, so we have shown that xn does indeed convergeto x as n → ∞. Hence ℓ1 is complete. ⊓⊔

In summary, some metric spaces have the property that every Cauchysequence in the space does converge to an element of the space. Since we cantest for Cauchyness without having the limit vector x in hand, this is oftenvery useful. We give such spaces the following name.

Definition 1.2.7 (Complete Metric Space). Let X be a metric space. Ifevery Cauchy sequence in X converges to an element of X, then we say thatX is complete. ♦

Theorem 1.2.6 shows that ℓ1 is a complete metric space. More precisely,it is complete with respect to the metric d(x, y) = ‖x − y‖1. If we imposea different metric on ℓ1, then it might not be the case that ℓ1 is completewith respect to that new metric. In particular, Problem 3.9.10 defines anothermetric on ℓ1 and shows that ℓ1 is not complete with respect to that metric.

Page 11: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.3 Topology in Metric Spaces 7

Remark 1.2.8. The reader should be aware that the term “complete” is heav-ily overused and has a number of distinct mathematical meanings. In par-ticular, the notion of a complete space as given in Definition 1.2.7 is quitedifferent from the notion of a complete sequence that will be introduced inDefinition 2.7.1. ♦

1.3 Topology in Metric Spaces

Given a metric space X, we make the following definitions.

• Given x ∈ X and r > 0, the open ball in X centered at x with radius r isthe set

Br(x) ={

y ∈ X : d(x, y) < r}

.

• A set U ⊆ X is open if for each x ∈ U there exists an r > 0 such thatBr(x) ⊆ U. Equivalently, U is open if and only if U can be written as aunion of open balls.

• The topology of X is the collection of all open subsets of X.

• The interior of a set E ⊆ X is the largest open set E◦ that is containedin E. Explicitly, E◦ =

S

{

U : U is open and U ⊆ E}

.

• A set E ⊆ X is closed if its complement X\E is open.

• A set E ⊆ X is bounded if it is contained in some open ball, i.e., thereexists some x ∈ X and r > 0 such that E ⊆ Br(x).

• A point x ∈ X is an accumulation point of a set E ⊆ X if there existxn ∈ E with all xn 6= x such that xn → x.

• A point x ∈ X is a boundary point of a set E ⊆ X if for every r > 0 wehave both Br(x) ∩ E 6= ∅ and Br(x) ∩ EC 6= ∅. The set of all boundarypoints of E is called the boundary of E, and it is denoted by ∂E.

• The closure of a set E ⊆ X is the smallest closed set E that contains E.Explicitly, E =

T

{

F : F is closed and E ⊆ F}

.

• A set E ⊆ X is dense in X if E = X.

• We say that X is separable if there exists a countable subset E that isdense in X.

The following lemma gives useful equivalent reformulations of some of thenotions defined above.

Lemma 1.3.1. If E is a subset of a metric space X, then the following state-ments hold.

Page 12: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

8 1 Metric Spaces

(a) E is closed if and only if

xn ∈ E and xn → x ∈ X =⇒ x ∈ E.

(b) The closure of E satisfies

E = ES

{y ∈ X : y is an accumulation point of E}

={

y ∈ X : there exist xn ∈ E such that xn → y}

. (1.5)

(c) E is dense in X if and only if for every point x ∈ X there exist xn ∈ Esuch that xn → x. ♦

Proof. (a) This statement can be deduced from part (b), or it can be proveddirectly. We assign the proof of statement (a) as Problem 1.6.11.

(b) Let F be the union of E and the accumulation points of E, i.e.,

F = ES

{y ∈ X : y is an accumulation point of E},

and let G be the set of all limits of points of E, i.e.,

G ={

y ∈ X : there exist xn ∈ E such that xn → y}

.

We must prove that F = E = G.To show that E ⊆ F, fix any point x ∈ FC = X \F. Then x /∈ E and

x is not an accumulation point of E. If every open ball Br(x) contained apoint y ∈ E, then x would be an accumulation point of E (because y cannotequal x). Therefore there must exist some ball Br(x) that contains no pointsof E. We claim Br(x) contains no accumulation points of E either. To seethis, suppose that y ∈ Br(x) was an accumulation point of E. Then therewould exist points xn ∈ E such that xn → y. By taking n large enough itfollows that xn ∈ Br(y), which is a contradiction. Thus Br(x) contains nopoints of E and no accumulation points of E, so Br(x) ⊆ FC. This showsthat FC is open, and therefore F is closed. Since E ⊆ F and E is the smallestclosed set that contains E, we conclude that E ⊆ F.

Next, we will prove that F ⊆ E by showing that E C ⊆ FC. Since E C isopen, if x belongs to this set then there exists some r > 0 such that

Br(x) ⊆ E C.

Hence Br(x) contains no points of E, so x cannot be an accumulation pointof E. We also have x /∈ E since x /∈ E, so it follows that x /∈ F.

The above work shows that E = F. Now choose any point y ∈ G. If y ∈ E,then y ∈ F since E ⊆ F. So, suppose that y ∈ G\E. Then, by definitionof G, there exist points xn ∈ E such that xn → y. But xn 6= y for any n sincexn ∈ E and y /∈ G, so y is an accumulation point of E. Therefore y ∈ F.Thus G ⊆ F.

Page 13: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.4 Compact Sets in Metric Spaces 9

Finally, suppose that y ∈ F. If y ∈ E, then y is the limit of the sequence{y, y, y, . . . }, so y ∈ G. On the other hand, if y ∈ F \E, then y must be anaccumulation point of E. Therefore there exist points xn ∈ E, with xn 6= yfor every n, such that xn → y. This shows that y is a limit of points from E,so y ∈ G. Therefore F ⊆ G.

(c) This follows from equation (1.5) and the fact that E is dense in X ifand only if X = E. ⊓⊔

Restating parts of Lemma 1.3.1, we see that:

• a set E is closed if and only if it contains every limit of points from E,

• the closure of a set E is the set of all limits of points from E, and

• a set E is dense in X if and only if every point in X is a limit of pointsfrom E.

For example, the set of rationals Q is not a closed subset of R because a limitof rational points need not be rational. The closure of Q is R because everypoint in R can be written as a limit of rational points. Similarly, Q is a densesubset of R because every real number can be written as a limit of rationalnumbers. Since Q is both countable and dense, this also shows that R is aseparable space. Problem 2.9.13 shows that the space ℓ1 defined in Example1.1.2 is separable.

1.4 Compact Sets in Metric Spaces

Now we define compact sets and discuss their properties.

Definition 1.4.1 (Compact Set). A subset K of a metric space X is com-pact if every covering of K by open sets has a finite subcovering. Statedprecisely, K is compact if it is the case that whenever

K ⊆⋃

i∈I

Ui,

where {Ui}i∈I is any collection of open subsets of X, there exist finitely manyi1, . . . , iN ∈ I such that

K ⊆N⋃

k=1

Uik. ♦

First we show that every compact subset of a metric space is both closedand bounded.

Lemma 1.4.2. If K is a compact subset of a metric space X, then K isclosed and there exists some open ball Br(x) such that K ⊆ Br(x).

Page 14: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

10 1 Metric Spaces

Proof. Suppose that K is compact, and fix any particular point x ∈ X. Thenthe union of the open balls Bn(x) over n ∈ N is all of X, so {Bn(x)}n∈N anopen cover of K. This cover must have a finite subcover, and since the ballsare nested it follows that K ⊆ Bn(x) for some single n.

It remains to show that K is closed. If K = X then we are done, soassume that K 6= X. Fix any point y in KC = X\K. Then given any x ∈ Kwe have x 6= y, so by the Hausdorff property stated in Problem 1.6.5, thereexist disjoint open sets Ux, Vx such that x ∈ Ux and y ∈ Vx. The collection{Ux}x∈K is an open cover of K, so it must contain some finite subcover, say

K ⊆ Ux1∪ · · · ∪ UxN

. (1.6)

Each Vxjis disjoint from Uxj

, so it follows from equation (1.6) that the set

V = Vx1∩ · · · ∩ VxN

is entirely contained in the complement of K. Thus, V is an open set thatsatisfies

y ∈ V ⊆ KC.

This shows that KC is open, so we conclude that K is closed. ⊓⊔

If X is a finite-dimensional normed space, then a subset of X is compactif and only if it closed and bounded (for a proof, see [Kre78, Thm. 2.5-3]or [Heil16]). However, every infinite-dimensional normed space contains aset that is closed and bounded but not compact. In fact, if X is an infinite-dimensional normed space then the closed unit ball in X is closed andbounded but not compact. Problem 2.9.12 asks for a proof of this fact for thespace ℓ1; for a proof in the general setting see [Heil11, Exercise 1.44].

We will give several equivalent reformulations of compactness for subsetsof metric spaces in terms of the following concepts.

Definition 1.4.3. Let E be a subset of a metric space X.

(a) E is sequentially compact if every sequence {xn}n∈N of points from Econtains a convergent subsequence {xnk

}k∈N whose limit belongs to E.

(b) E is totally bounded if given any r > 0 we can cover E by finitely manyopen balls of radius r. That is, for each r > 0 there must exist finitelymany points x1, . . . , xN ∈ X such that

E ⊆N⋃

k=1

Br(xk). ♦

We will need the following lemma.

Lemma 1.4.4. Let E be a sequentially compact subset of a metric space X.If {Ui}i∈I is an open cover of E, then there exists a number δ > 0 such thatif B is an open ball of radius δ that intersects E, then there is an i ∈ I suchthat B ⊆ Ui.

Page 15: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.4 Compact Sets in Metric Spaces 11

Proof. Let {Ui}i∈I be an open cover of E. We want to prove that there is aδ > 0 such that

Br(x) ∩ E 6= ∅ =⇒ Br(x) ⊆ Ui for some i ∈ I.

Suppose that no δ > 0 has this property. Then for each positive integer n,there must exist some open ball with radius 1

n that intersects E but is not

contained in any set Ui. Call this open ball Gn. For each n, choose a pointxn ∈ Gn ∩ E. Then since {xn}n∈N is a sequence in E and E is sequentiallycompact, there must be a subsequence {xnk

}k∈N that converges to a point

x ∈ E. Since {Ui}i∈I is a cover of E, we must have x ∈ Ui for some i ∈ I,and since Ui is open there must exist some r > 0 such that Br(x) ⊆ Ui. Nowchoose k large enough that we have both

1

nk<

r

3and d

(

x, xnk

)

<r

3.

Keeping in mind the facts that Gnkcontains xnk

, Gnkis an open ball with

radius 1/nk, the distance from x to xnkis less than r/3, and Br(x) has

radius r, it follows that Gnk⊆ Br(x) ⊆ Ui, which is a contradiction. ⊓⊔

Now we prove some reformulations of compactness that hold for subsetsof metric spaces.

Theorem 1.4.5. If K is a subset of a complete metric space X, then thefollowing statements are equivalent.

(a) K is compact.

(b) K is sequentially compact.

(c) K is totally bounded and every Cauchy sequence of points from K con-verges to a point in K.

Proof. (a) ⇒ (b). Suppose that E is not sequentially compact. Then thereexists a sequence {xn}n∈N in E that has no subsequence that converges toan element of E. Choose any point x ∈ E. If every open ball centered at xcontains infinitely many of the points xn, then by considering radii r = 1

kwe can construct a subsequence {xnk

}k∈N that converges to x. This is acontradiction, so there must exist some open ball centered at x that containsonly finitely many xn. If we call this ball Bx, then {Bx}x∈E is an open coverof E that contains no finite subcover. Consequently E is not compact.

(b) ⇒ (c). Suppose that E is sequentially compact, and suppose that{xn}n∈N is a Cauchy sequence in E. Then since E is sequentially compact,there is a subsequence {xnk

}k∈N that converges to some point x ∈ E. Hence{xn}n∈N is Cauchy and has a convergent subsequence. Appealing to Problem1.6.8, this implies that xn → x. Therefore E is complete.

Suppose that E is not totally bounded. Then there is a radius r > 0 suchthat E cannot be covered by finitely many open balls of radius r centered

Page 16: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

12 1 Metric Spaces

at points of X. Choose any point x1 ∈ E. Since E cannot be covered bya single r-ball, E cannot be a subset of Br(x1). Hence there exists a pointx2 ∈ E\Br(x1). In particular, d(x2, x1) ≥ r. But E cannot be covered by twor-balls, so there must exist a point x3 that belongs to E \

(

Br(x1)∪Br(x2))

.In particular, we have both d(x3, x1) ≥ r and d(x3, x2) ≥ r. Continuing inthis way we obtain a sequence of points {xn}n∈N in E that has no convergentsubsequence, which is a contradiction.

(c) ⇒ (b). Assume that E is complete and totally bounded, and let{xn}n∈N be any sequence of points in E. Since E can be covered by finitelymany open balls of radius 1

2 , at least one of those balls must contain infinitely

many of the points xn. That is, there is some infinite subsequence {x(1)n }n∈N

of {xn}n∈N that is contained in an open ball of radius 12 . The Triangle In-

equality therefore implies that

∀m,n ∈ N, d(

x(1)m , x(1)

n

)

< 1.

Similarly, since E can be covered by finitely many open balls of radius 14 ,

there is some subsequence {x(2)n }n∈N of {x

(1)n }n∈N such that

∀m,n ∈ N, d(

x(2)m , x(2)

n

)

<1

2.

Continuing by induction, for each k > 1 we find a subsequence {x(k)n }n∈N of

{x(k−1)n }n∈N such that d

(

x(k)m , x

(k)n

)

< 1k for all m, n ∈ N.

Now consider the diagonal subsequence {x(k)k }k∈N. Given ε > 0, let n be

large enough that 1n < ε. If j ≥ k > n, then x

(j)j is one element of the

sequence {x(k)n }n∈N (why?), say x

(j)j = x

(k)n . Hence

d(

x(j)j , x

(k)k

)

= d(

x(k)n , x

(k)k

)

<1

k<

1

n< ε.

Thus {x(k)k }k∈N is a Cauchy subsequence of the original sequence {xn}n∈N.

Since K is complete, this subsequence must converge to some element of E.Hence K is sequentially compact.

(b) ⇒ (a). Assume that E is sequentially compact. Since we have alreadyproved that statement (b) implies statement (c), we know that E is completeand totally bounded.

Suppose that {Ui}i∈J is any open cover of E. By Lemma 1.4.4, thereexists a δ > 0 such that if B is an open ball of radius δ that intersects E,then there is an i ∈ J such that B ⊆ Ui. However, E is totally bounded, sowe can cover E by finitely many open balls of radius δ. Each of these balls iscontained in some Ui, so E is covered by finitely many Ui. ⊓⊔

Page 17: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.5 Continuity for Functions on Metric Spaces 13

1.5 Continuity for Functions on Metric Spaces

Here is the abstract definition of continuity for functions on metric spaces.

Definition 1.5.1 (Continuous Function). Let X and Y be metric spaces.We say that a function f : X → Y is continuous if given any open set V ⊆ Y,its inverse image f−1(V ) is an open subset of X. ♦

By applying the definition of continuity, the following lemma shows thatthe direct image of a compact set under a continuous function is compact.

Lemma 1.5.2. Let X and Y be metric spaces. If f : X → Y is continuousand K is a compact subset of X, then f(K) is a compact subset of Y.

Proof. Let {Vi}i∈J be any open cover of f(K). Each set Ui = f−1(Vi) is open,and {Ui}i∈J is an open cover of K. Since K is compact, this cover must havea finite subcover {Ui1 , . . . , UiN

}. But then {Vi1 , . . . , ViN} is a finite subcover

of f(K), so we conclude that f(K) is compact. ⊓⊔

The next lemma gives a useful reformulation of continuity for functions onmetric spaces in terms of preservation of limits.

Lemma 1.5.3. Let X be a metric space with metric dX , and let Y be a metricspace with metric dY . If f : X → Y, then the following three statements areequivalent.

(a) f is continuous.

(b) If x is any point in X, then for every ε > 0 there exists a δ > 0 such that

dX(x, y) < δ =⇒ dY

(

f(x), f(y))

< ε.

(c) Given any point x ∈ X and any sequence {xn}n∈N in X,

xn → x in X =⇒ f(xn) → f(x) in Y. ♦

Proof. For this proof we let BXr (x) and BY

s (y) denote open balls in X and Y,respectively.

(a) ⇒ (b). Suppose that f is continuous, and choose any point x ∈ Xand any ε > 0. Then the ball V = BY

ε

(

f(x))

is an open subset of Y, soU = f−1(V ) must be an open subset of X. As x ∈ U, there exists some δ > 0such that BX

δ (x) ⊆ U. If y ∈ X is any point that satisfies dX(x, y) < δ, thenwe have y ∈ BX

δ (x) ⊆ U, and therefore

f(y) ∈ f(U) ⊆ V = BYε

(

f(x))

.

Consequently dY

(

f(x), f(y))

< ε.

(b) ⇒ (c). Assume that statement (b) holds, choose any point x ∈ X,and suppose that points xn ∈ X are such that xn → x. Fix any ε > 0, and

Page 18: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

14 1 Metric Spaces

let δ > 0 be the number whose existence is given by statement (b). Sincexn → x, there must exist some N > 0 such that dX(x, xn) < δ for all n > N.Statement (b) therefore implies that dY

(

f(x), f(xn))

< ε for all n > N, sowe conclude that f(xn) → f(x) in Y.

(c) ⇒ (a). Suppose that statement (c) holds, and let V be any open subsetof Y. Suppose that f−1(V ) was not open. Then f−1(V ) cannot be the emptyset, so there must exist some point x ∈ f−1(V ). By definition, this meansthat y = f(x) ∈ V. Since V is open, there exists some ε > 0 such thatBY

ε (y) ⊆ V. But f−1(V ) is not open, so for no n ∈ N can the ball BX1/n(x) be

entirely contained in f−1(V ). Hence for every n there is some xn ∈ BX1/n(x)

such that xn /∈ f−1(V ). Since dX(x, xn) < 1/n, we have xn → x. Applyingstatement (c), it follows that f(xn) → f(x). Therefore there must exist someN > 0 such that dY

(

f(x), f(xn))

< ε for all n > N. Hence for every n > Nwe have f(xn) ∈ BY

ε (y) ⊆ V, and therefore xn ∈ f−1(V ). This contradictsthe definition of xn, so we conclude that f−1(V ) must be open. Therefore fis continuous. ⊓⊔

The number δ that appears in statement (b) of Lemma 1.5.3 depends bothon the point x and the number ε. We say that a function f is uniformly con-tinuous if δ can be chosen independently of x. Here is the precise definition.

Definition 1.5.4 (Uniform Continuity). Let X be a metric space withmetric dX , and let Y be a metric space with metric dY . If E ⊆ X, then wesay that a function f : X → Y is uniformly continuous on E if for every ε > 0there exists a δ > 0 such that

∀x, y ∈ E, dX(x, y) < δ =⇒ dY

(

f(x), f(y))

< ε. ♦

The next lemma shows that a continuous function whose domain is acompact set is uniformly continuous on that set.

Lemma 1.5.5. Let X and Y be metric spaces. If K is a compact subset of Xand f : K → Y is continuous, then f is uniformly continuous on K.

Proof. For this proof, let dX and dY denote the metrics on X and Y, and letBX

r (x) and BYs (y) denote open balls in X and Y, respectively.

Suppose that f : K → Y is continuous, and fix ε > 0. For each pointz ∈ Y, let Uz = f−1(BY

ε (z)). Then {Uz}z∈Y is an open cover of K. Since Kis compact, it is sequentially compact. Therefore Lemma 1.4.4 implies thatthere exists a number δ > 0 such that if B is any open ball of radius δ in Xthat intersects K, then B ⊆ Uz for some z ∈ Y.

Now choose any points x, y ∈ K with dX(x, y) < δ. Then x, y ∈ BXδ (x),

and BXδ (x) must be contained in some set Uz, so

f(x), f(y) ∈ f(BXδ (x)) ⊆ f(Uz) ⊆ BY

ε (z).

Therefore dY (f(x), f(y)) < 2ε, so f is uniformly continuous. ⊓⊔

Page 19: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.6 Urysohn’s Lemma 15

1.6 Urysohn’s Lemma

Urysohn’s Lemma is a basic result about the existence of continuous func-tions. We will state this theorem for functions on metric spaces, althoughit can be generalized to functions on normal topological spaces (for precisestatements and proofs, see texts such as [Mun75] or [ST76]).

To prove Urysohn’s Lemma for functions on a metric space, we will use thefollowing lemma, which states that the distance from a point x to a subset Eof a metric space is a continuous function of x.

Lemma 1.6.1. Let X be a metric space. If E is a nonempty subset of X,then the function f : X → R defined by

f(x) = dist(x,E) = inf{

d(x, z) : z ∈ E}

, x ∈ X,

is uniformly continuous on X.

Proof. Fix ε > 0, and set δ = ε/2. Choose any points x, y ∈ X such thatd(x, y) < δ. By definition of the distance function, there exist points a, b ∈ Esuch that

d(x, a) < dist(x,E) + δ and d(y, b) < dist(y,E) + δ.

Consequently,

f(y) = dist(y,E) ≤ d(y, a)

≤ d(y, x) + d(x, a)

< δ +(

dist(x,E) + δ)

= f(x) + ε.

Interchanging the roles of x and y, we similarly obtain f(x) < f(y) + ε.Therefore |f(x) − f(y)| < ε whenever d(x, y) < δ. This tells us that f isuniformly continuous on X. ⊓⊔

Now we prove Urysohn’s Lemma, which states that if E and F are disjointclosed sets, then we can find a continuous function that is identically 0 on Eand identically 1 on F. In this sense, E and F can be “separated” by acontinuous function.

Theorem 1.6.2 (Urysohn’s Lemma). If E, F are disjoint closed subsetsof a metric space X, then there exists a continuous function θ : X → R suchthat

(a) 0 ≤ θ ≤ 1 on X,

(b) θ = 0 on E, and

(c) θ = 1 on F.

Page 20: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

16 1 Metric Spaces

Proof. Since E is closed, if x /∈ E then dist(x,E) > 0. Lemma 1.6.1 tells usdist(x,E) and dist(x, F ) are each continuous functions of x. It follows thatthe function

θ(x) =dist(x,E)

dist(x,E) + dist(x, F )

has the required properties. ⊓⊔

We can sometimes improve on Urysohn’s Lemma Theorem when dealingwith particular concrete domains. For example, the following result says thatif we take X = R and fix a compact set K ⊆ R, then we can find an infinitelydifferentiable, compactly supported function that is identically 1 on K.

Theorem 1.6.3. If K ⊆ R is compact and U ⊇ K is open, then there existsa function f ∈ C∞

c (R) such that

(a) 0 ≤ f ≤ 1 everywhere,

(b) f = 1 on K, and

(c) supp(f) ⊆ U. ♦

One way to prove Theorem 1.6.3 is by combining Urysohn’s Lemma withthe operation of convolution. For details on convolution we refer to texts suchas [DM72], [Ben97], or [Kat04]. For comparison, Problem 1.6.20 presentsa less ambitious result. Specifically, it shows that there exists an infinitelydifferentiable function g that is identically zero outside of the finite interval[0, 1] and nonzero on the interior of this interval. Problems 1.6.21 and 1.6.22give some related constructions.

Problems

1.6.4. Given a set X, show that

d(x, y) =

{

1, x 6= y,

0, x = y,

is a metric on X. This is called the discrete metric on X.

1.6.5. Prove that every metric space X is Hausdorff, i.e., if x 6= y are twodistinct elements of X, then there exist disjoint open sets U, V such thatx ∈ U and y ∈ V.

1.6.6. Let X be a metric space X. Prove that the following inequality (calledthe Reverse Triangle Inequality) holds for all x, y, z ∈ X:

∣d(x, z) − d(y, z)∣

∣ ≤ d(x, y).

Page 21: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.6 Urysohn’s Lemma 17

1.6.7. Let X be a metric space. Prove that the limit of any convergent se-quence in X is unique, i.e., if xn → y and xn → z then y = z.

1.6.8. Suppose that {xn}n∈N is a Cauchy sequence in a metric space X, andsuppose there exists a subsequence {xnk

}k∈N that converges to x ∈ X, i.e.,xnk

→ x as k → ∞. Prove that xn → x as n → ∞.

1.6.9. Let {xn}n∈N be a sequence in a metric space X, and fix x ∈ X. Supposethat every subsequence {yn}n∈N of {xn}n∈N has a subsequence {zn}n∈N of{yn}n∈N such that zn → x. Prove that xn → x.

1.6.10. Let A be a subset of a metric space X.

(a) Prove that A is open if and only if A = A◦.

(b) Prove that A is closed if and only if A = A.

1.6.11. Prove part (a) of Lemma 1.3.1.

1.6.12. Let X be a complete metric space, and let A be a subset of X. Provethat the following two statements are equivalent.

(a) A is closed.

(b) A is complete, i.e., every Cauchy sequence of points in A converges toa point of A.

1.6.13. (a) Let Q be the “open first quadrant” in ℓ1, i.e.,

Q ={

x = (xk)k∈N ∈ ℓ1 : xk > 0 for every k}

.

Prove that R is not an open subset of ℓ1.

(b) Let R be the “closed first quadrant” in ℓ1, i.e.,

R ={

x = (xk)k∈N ∈ ℓ1 : xk ≥ 0 for every k}

.

Prove that R is a closed subset of ℓ1.

1.6.14. Let X be a metric space. The diameter of a subset S of X isdiam(S) = sup

{

d(x, y) : x, y ∈ S}

.

(a) Suppose that X is complete and F1 ⊇ F2 ⊇ · · · is a nested decreasingsequence of closed nonempty subsets of X such that diam(Fn) → 0. Provethat there exists some x ∈ X such that

T

Fn = {x}.

(b) Show by example that the conclusion of part (a) can fail if X is notcomplete.

1.6.15. Let E be a subset of a metric space X. Given x ∈ X, prove that thefollowing four statements are equivalent.

(a) x is an accumulation point of E.

(b) If U is an open set that contains x, then (E ∩U) \ {x} 6= ∅, i.e., thereexists a point y ∈ E ∩ U such that y 6= x.

Page 22: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

18 1 Metric Spaces

(c) If r > 0, then exists a point y ∈ E such that 0 < d(x, y) < r.

(d) Every open set U that contains x also contains infinitely many pointsof E.

1.6.16. Let X be a metric space.

(a) Given A, B ⊆ X, prove that A ∪ B = A ∪ B.

(b) Let I be an arbitrary index set. Given sets Ei ⊆ X for i ∈ I, prove that

i∈I

Ei ⊆⋂

i∈I

Ei.

Show by example that equality can fail.

1.6.17. Let K be a compact subset of a metric space X, and let E be a subsetof K. Prove that E is closed if and only if E is compact.

1.6.18. Let K be a compact subset of a metric space X, and assume thatf : K → R is continuous. Prove that f achieves its maximum and its minimumon K, i.e., there exist points x, y ∈ K such that

f(x) = inf{

f(t) : t ∈ K}

and f(y) = sup{

f(t) : t ∈ K}

.

1.6.19. (a) Let X and Y be metric spaces, and suppose that f : X → Yis uniformly continuous. Prove that f maps Cauchy sequences to Cauchysequences, i.e.,

{xn}n∈N is Cauchy in X =⇒ {f(xn)}n∈N is Cauchy in Y.

(b) Show by example that part (a) can fail if we only assume that f iscontinuous instead of uniformly continuous. (Contrast this with Lemma 1.5.3,which shows that every continuous function must map convergent sequencesto convergent sequences.)

1.6.20. Define

γ(x) = e−1/x χ[0,∞)(x) =

{

e−1/x, x > 0,

0, x ≤ 0,

and

β(x) = γ(

1 − x2)

=

{

e−1/(x2−1), −1 < x < 1,

0, |x| ≥ 1.

Prove the following statements.

(a) γ(x) = 0 for all x ≤ 0, and γ(x) > 0 for all x > 0.

(b) For each n ∈ N, there exists a polynomial pn of degree n− 1 such that

γ(n)(x) =pn(x)

x2nγ(x).

Page 23: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

1.6 Urysohn’s Lemma 19

(c) γ ∈ C∞(R) and γ(n)(0) = 0 for every n ≥ 0.

(d) β ∈ C∞c (R), β(x) > 0 for |x| < 1, and β(x) = 0 for |x| ≥ 1.

1.6.21. Let f be the piecewise linear function pictured in Figure 1.1. Showthat the function

g(x) =

∫ x

−3

f(t) dt, x ∈ R,

has the following properties.

f

-4 -2 2 4

-1

-0.5

0.5

1

Fig. 1.1 The function f from Problem 1.6.21.

(a) 0 ≤ g(x) ≤ 1 for every x ∈ R.

(b) g(x) = 1 for |x| ≤ 1.

(c) g(x) = 0 for |x| ≥ 3.

(d) g ∈ C1c (R), but g /∈ C2

c (R).

1.6.22. Prove that there exists an infinitely differentiable function θ thatsatisfies:

(a) 0 ≤ θ(x) ≤ 1 for every x ∈ R,

(b) θ(x) = 1 for all |x| ≤ 1,

(c) θ(x) = 0 for all |x| ≥ 3.

Page 24: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric
Page 25: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Chapter 2

Norms and Banach Spaces

While a metric provides us with a notion of the distance between pointsin a space, a norm gives us a notion of the length of an individual vector. Anorm can only be defined on a vector space, while a metric can be defined onarbitrary sets.

2.1 The Definition of a Norm

Here is the definition of a norm, and the slightly weaker notion of a seminorm.

Definition 2.1.1 (Seminorms and Norms). Let X be a vector space overthe real field R or the complex field C. A seminorm on X is a function‖ · ‖ : X → R such that for all vectors x, y ∈ X and all scalars c we have:

(a) Nonnegativity: ‖x‖ ≥ 0,

(b) Homogeneity: ‖cx‖ = |c| ‖x‖, and

(c) The Triangle Inequality: ‖x + y‖ ≤ ‖x‖ + ‖y‖.

A seminorm is a norm if we also have:

(d) Uniqueness: ‖x‖ = 0 if and only if x = 0.

A vector space X together with a norm ‖ · ‖ is called a normed vectorspace, a normed linear space, or simply a normed space. ♦

We refer to the number ‖x‖ as the length of a vector x, and we say that‖x − y‖ is the distance between the vectors x and y. A vector x that haslength 1 is called a unit vector, or is said to be normalized.

We usually use ‖ · ‖ to denote a norm or seminorm, sometimes with sub-scripts to distinguish among different norms (for example, ‖ · ‖a and ‖ · ‖b).Other common symbols for norms or seminorms are | · |, ||| · |||, or ρ(·). Theabsolute value function |x| is a norm on the real line R and on the complexplane C. Likewise, the Euclidean norm is a norm on Rd and Cd.

c©2014 Christopher Heil21

Page 26: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

22 2 Norms and Banach Spaces

Example 2.1.2. The space ℓ1 introduced in Example 1.1.2 is a vector space,and it is straightforward to check that the “ℓ1-norm”

‖x‖1 =

∞∑

k=1

|xk|, x = (xk)k∈N ∈ ℓ1,

defined in equation (1.1) is indeed a norm on ℓ1. Therefore ℓ1 is a normedspace. ♦

2.2 The Induced Metric

If X is a normed space, then it follows directly from the definition of a normthat

d(x, y) = ‖x − y‖, x, y ∈ X,

defines a metric on X. This is called the metric on X induced from ‖ · ‖, orsimply the induced metric on X.

Consequently, whenever we are given a normed space X, we have a metricon X as well as a norm, and this metric is the induced metric. Therefore alldefinitions made for metric spaces apply to normed spaces, using the inducednorm d(x, y) = ‖x−y‖. For example, convergence in a normed space is definedby

xn → x ⇐⇒ limn→∞

‖x − xn‖ = 0.

It may be possible to place a metric on X other than the induced metric,but unless we explicitly state otherwise, all metric-related statements on anormed space are taken with respect to the induced metric.

2.3 Properties of Norms

Here are some useful properties of norms (we assign the proof as Problem2.9.6).

Lemma 2.3.1. If X is a normed space, then the following statements hold.

(a) Reverse Triangle Inequality:∣

∣‖x‖ − ‖y‖∣

∣ ≤ ‖x − y‖.

(b) Convergent implies Cauchy: If xn → x, then {xn}n∈N is Cauchy.

(c) Boundedness of Cauchy sequences: If {xn}n∈N is a Cauchy sequence, thensup ‖xn‖ < ∞.

(d) Continuity of the norm: If xn → x, then ‖xn‖ → ‖x‖.

(e) Continuity of vector addition: If xn → x and yn → y, then xn+yn → x+y.

Page 27: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.5 Banach Spaces 23

(f) Continuity of scalar multiplication: If xn → x and cn → c, thencnxn → cx. ♦

2.4 Convexity

If x and y are vectors in a vector space X, then the line segment joining xto y is the set of all points of the form tx + (1 − t)y where 0 ≤ t ≤ 1.

A subset K of a vector space is convex if given any two points x, y ∈ K,the line segment joining x to y is entirely contained within K. That is, K isconvex if

x, y ∈ K, 0 ≤ t ≤ 1 =⇒ tx + (1 − t)y ∈ K.

If X is a normed space, then every open ball Br(x) in X is convex (this isProblem 2.9.7). However, if X is a vector space that is merely a metric spacebut not a normed space then open balls in X need not be convex.

2.5 Banach Spaces

Every convergent sequence in a normed vector space must be Cauchy, butthe converse does not hold in general. In some normed spaces it is true thatevery Cauchy sequence in the space is convergent. We give spaces that havethis property the following name.

Definition 2.5.1 (Banach Space). A normed space X is a Banach spaceif it is complete, i.e., if every Cauchy sequence in X converges to an elementof X. ♦

The terms “Banach space” and “complete normed space” are interchange-able, and we will use whichever is more convenient in a given context.

The set of real numbers R is complete with respect to absolute value (usingreal scalars), and likewise the complex plane C is complete with respect toabsolute value (using complex scalars). More generally, Rd and Cd are Banachspaces with respect to the Euclidean norm. In fact, every finite-dimensionalvector space V is complete with respect to any norm that we place on V (fora proof, see [Con90, Sec. III.3] or [Heil16]).

We proved in Theorem 1.2.6 that the space ℓ1 is complete. Therefore ℓ1

is an example of an infinite-dimensional Banach space. One example of aninfinite-dimensional normed space that is not a Banach space is the space c00

studied in Example 1.2.5. A space that is complete with respect to one normmay be incomplete if replace that norm with another norm. For example, ℓ1

is no longer complete if we replace the norm ‖ · ‖1 with the “ℓ2-norm” (seeProblem 3.9.10).

Page 28: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

24 2 Norms and Banach Spaces

2.6 Infinite Series in Normed Spaces

An infinite series is a limit of the partial sums of the series. Hence, thedefinition of an infinite series involves both addition and limits, each of whichare defined in normed spaces. The precise definition of an infinite series in anormed space is as follows.

Definition 2.6.1 (Convergent Series). Let {xn}n∈N be a sequence in a

normed space X. We say that the series∑∞

n=1 xn converges and equals x ∈ X

if the partial sums sN =∑N

n=1 xn converge to x, i.e., if

limN→∞

‖x − sN‖ = limN→∞

x −

N∑

n=1

xn

= 0.

In this case, we write x =∑∞

n=1 xn, and we also use the shorthands x =∑

xn

or x =∑

n xn. ♦

In order for an infinite series to converge in X, the norm of the differencebetween f and the partial sum sN must converge to zero. If we wish toemphasize which norm we are referring to, we may write that x =

xn

converges with respect to ‖ · ‖, or we may say that x =∑

xn converges in X.Here is another type of convergence notion for series.

Definition 2.6.2. Let {xn}n∈N be a sequence in a normed space X. We saythat the series

∑∞n=1 xn is absolutely convergent if

∞∑

n=1

‖xn‖ < ∞. ♦

Note that Definition 2.6.2 does not imply that an absolutely convergentseries will converge in the sense of Definition 2.6.1. We will prove in thenext theorem that if X is complete then every absolutely convergent seriesin X converges, while if X is not complete then there will exist some vectorsxn ∈ X such that

‖xn‖ < ∞ yet∑

xn does not converge.

Theorem 2.6.3. If X is a normed space, then the following two statementsare equivalent.

(a) X is complete (i.e., X is a Banach space).

(b) Every absolutely convergent series in X converges in X. That is, if{xn}n∈N is a sequence in X and

‖xn‖ < ∞, then the series∑

xn

converges in X.

Proof. (a) ⇒ (b). Assume that X is complete, and suppose that∑

‖xn‖ < ∞.Set

Page 29: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.6 Infinite Series in Normed Spaces 25

sN =

N∑

n=1

xn and tN =

N∑

n=1

‖xn‖.

Given any N > M we have

‖sN − sM‖ =

N∑

n=M+1

xn

≤N

n=M+1

‖xn‖ = |tN − tM |.

Since {tN}N∈N is a Cauchy sequence of scalars, this implies that {sN}N∈N isa Cauchy sequence of vectors in X, and hence converges. By definition, thismeans that

∑∞n=1 xn converges in X.

(b) ⇒ (a). Suppose that every absolutely convergent series in X is con-vergent. Let {xn}n∈N be a Cauchy sequence in X. Applying Problem 2.9.9,there exists a subsequence {xnk

}k∈N such that

∀ k ∈ N, ‖xnk+1− xnk

‖ < 2−k.

The series∑∞

k=1 (xnk+1− xnk

) is absolutely convergent, because

∞∑

k=1

‖xnk+1− xnk

‖ ≤

∞∑

k=1

2−k = 1 < ∞.

Therefore the series∑∞

k=1 (xnk+1− xnk

) converges in X, so we can set

x =∞∑

k=1

(xnk+1− xnk

).

By definition, this means that the partial sums

sM =

M∑

k=1

(xnk+1− xnk

) = xnM+1− xn1

converge to x as M → ∞. Setting y = x + xn1, it follows that

xnM+1= sM + xn1

→ x + xn1= y as M → ∞.

Reindexing (replace M + 1 by k), we conclude that xnk→ y as k → ∞.

Thus {xn}n∈N is a Cauchy sequence that has a subsequence {xnk}k∈N

that converges to the vector y. Appealing to Problem 1.6.8, this implies thatxn → y. Hence every Cauchy sequence in X converges, so X is complete. ⊓⊔

The ordering of the vectors in a series may be important. If we reorder aseries, or in other words consider a new series

∑∞n=1 xσ(n) where σ : N → N is

a bijection, there is no guarantee that this reordered series will still converge.If

∑∞n=1 xσ(n) does converge for every bijection σ, then we say that the series

Page 30: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

26 2 Norms and Banach Spaces

∑∞n=1 xn converges unconditionally. A series that converges but does not

converge unconditionally is said to be conditionally convergent.For series of scalars, unconditional and absolute convergence are equivalent

(for one proof of the following result, see [Heil11, Lemma 3.3]).

Lemma 2.6.4. If (cn)n∈N is a sequence of scalars, then

∞∑

n=1

|cn| < ∞ ⇐⇒

∞∑

n=1

cn converges unconditionally. ♦

More generally, if X is a finite-dimensional normed space, then an infi-nite series converges absolutely if and only if it converges unconditionally.The proof of the following theorem on infinite-dimensional spaces is non-trivial. This result is a consequence of the Dvoretzky–Rogers Theorem (seeTheorem 3.33 in [Heil11]).

Theorem 2.6.5. If X is an infinite dimensional Banach space, then thereexist vectors xn ∈ X such that the infinite series

xn converges uncondi-tionally but not absolutely. ♦

2.7 Span and Closed Span

If A is a subset of a vector space X, then the finite linear span of A, denotedby span(A), is the set of all possible finite linear combinations of elementsof A. If the scalar field is C, then

span(A) =

{ N∑

n=1

cnxn : N > 0, xn ∈ A, cn ∈ C

}

, (2.1)

and if the scalar field is R then we need only replace C by R on the precedingline. We often call span(A) the finite span, the linear span, or simply the spanof A. We say that A spans X if span(A) = X.

If X is a normed space, then we can take limits of elements, and considerthe closure of the span. The closure of span(A) is called the closed linear spanor simply the closed span of A. For compactness of notation, usually write

the closed span as span(A) instead of span(A).It follows from part (b) of Lemma 1.3.1 that the closed span consists of

all limits of elements of the span:

span(A) ={

y ∈ X : ∃ yn ∈ span(A) such that yn → y}

. (2.2)

Suppose that xn ∈ A, cn is a scalar, and x =∑

cnxn is a convergentinfinite series in X. Then the partial sums

Page 31: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.8 Hamel Bases and Schauder Bases 27

sN =

N∑

n=1

cnxn

belong to span(A) and sN → x, so it follows from equation (2.2) that xbelongs to span(A). As a consequence, we have the inclusion

{ ∞∑

n=1

cnxn : xn ∈ A, cn scalar,

∞∑

n=1

cnxn converges

}

⊆ span(A). (2.3)

However, equality need not hold in equation (2.3). A specific example of sucha situation is presented in Example 2.9.5.

According to the following definition, if the closed span of a sequence isthe entire space X, then we say that sequence is complete.

Definition 2.7.1 (Complete Sequence). Let {xn}n∈N be a sequence ofvectors in a normed vector space X. We say that the sequence {xn}n∈N iscomplete in X if span{xn}n∈N is dense in X, i.e., if

span{xn}n∈N = X.

Complete sequences are also known as total or fundamental sequences. ♦

If M is a subspace of a normed space X, then, by definition, M is finitedimensional if it has a finite Hamel basis. Equivalently, a subspace M isfinite dimensional if and only if we can write M = span{x1, . . . , xn} for somefinitely many vectors x1, . . . , xn ∈ X. The following result states that everyfinite-dimensional subspace of a normed space is closed. For a proof of thistheorem, see [Con90, Prop. III.3.3]. or [Heil16].

Theorem 2.7.2. If M is a finite-dimensional subspace of a normed space X,then M is a closed subset of X. ♦

2.8 Hamel Bases and Schauder Bases

A Hamel Basis is simply another name for the usual notion of a basis for avector space. We formalize this in the following definition.

Definition 2.8.1 (Hamel Basis). Let V be a vector space. A Hamel basis,vector space basis, or simply a basis for V is a set B ⊆ V such that

(a) B is linearly independent, and

(b) span(B) = V. ♦

The span of B was defined in equation (2.1) to be the set of all finite linearcombinations of vectors from B. Likewise, independence is defined in terms

Page 32: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

28 2 Norms and Banach Spaces

of finite linear combinations. Specifically, B is linearly independent if givenany finitely many distinct vectors x1, . . . , xN ∈ B we have

N∑

n=1

cnxn = 0 ⇐⇒ c1 = · · · = cN = 0.

Thus the definition of a Hamel basis, as given in Definition 2.8.1, is madeentirely in terms of finite linear combinations of vectors. This “vector space”notion of a basis is most useful when V is a finite-dimensional vector space.One consequence of the Baire Category Theorem is that any Hamel basis foran infinite-dimensional Banach space must be uncountable (for details, see[Heil11, Sec. 4.1]).

In an infinite-dimensional Banach space, instead of restricting to just fi-nite linear combinations, it is usually preferable to employ “infinite linearcombinations.” This idea leads us to the following definition of a Schauderbasis for a Banach space.

Definition 2.8.2 (Schauder Basis). Let X be a Banach space. A countablyinfinite sequence {xn}n∈N of elements of X is a Schauder basis for X if foreach vector x ∈ X there exist unique scalars cn(x) such that

x =∞∑

n=1

cn(x)xn, (2.4)

where this series converges in the norm of X. ♦

Recall from Definition 2.6.1 that equation (2.4) means that the partialsums of the series converge to x in the norm of X, i.e.,

limN→∞

x −

N∑

n=1

cn(x)xn

= 0.

Example 2.8.3. Let E = {δn}n∈N be the sequence of standard basis vectors,and let x be any element of the space ℓ1. Then we can write

x = (xn)n∈N = (x1, x2, . . . )

where the xn are scalars such that∑

|xn| < ∞. We claim that

x =∞∑

n=1

xnδn, (2.5)

with convergence of this series in ℓ1-norm. To see why, let

sN =N

n=1

xnδn = (x1, . . . , xN , 0, 0, . . . )

Page 33: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.9 The Space Cb(X) 29

be the Nth partial sum of the series. Then

x − sN = (0, . . . , 0, xN+1, xN+2, . . . ).

Therefore ‖x − sN‖1 =∑∞

n=N+1 |xn|, and since∑

|xn| < ∞, it follows that

limN→∞

‖x − sN‖1 = limN→∞

x −

N∑

n=1

xnδn

1

= limN→∞

∞∑

n=N+1

|xn| = 0.

This shows that the partial sums sN converge to x in ℓ1-norm, which provesthat equation (2.5) holds. Further, the only way to write x =

cnδn is withcn = xn for every n (consider Problem 2.9.11), so we conclude that E is aSchauder basis for the space ℓ1. ♦

The question of whether every separable Banach space has a Schauderbasis was a longstanding open problem known as the Basis Problem. It wasfinally shown by Enflo [Enf73] that there exist separable Banach spaces thathave no Schauder bases!

For an introduction to the theory of Schauder bases and related topics andgeneralizations, we refer to [Heil11].

2.9 The Space Cb(X)

Let X be a metric space, and let Cb(X) be the space of all bounded, con-tinuous functions f : X → C (entirely analogous statements if we prefer torestrict our attention to real-valued functions only). The uniform norm of afunction f ∈ Cb(X) is

‖f‖u = supx∈X

|f(x)|.

This is a norm on Cb(X).Convergence with respect to the norm ‖ · ‖u is called uniform convergence.

That is, if fn, f ∈ Cb(X) and

limn→∞

‖f − fn‖u = limn→∞

(

supx∈X

|f(x) − fn(x)|

)

= 0,

then we say that fn converges uniformly to f. Similarly, a sequence {fn}n∈N

that is Cauchy with respect to the uniform norm is said to be a uniformlyCauchy sequence.

Suppose that fn converges uniformly to f. Then, by definition of a supre-mum, for each individual point x we have

limn→∞

|f(x) − fn(x)| ≤ limn→∞

‖f − fn‖u = 0.

Page 34: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

30 2 Norms and Banach Spaces

Thus fn(x) → f(x) for each individual x. We have a name for this lattertype of convergence. Specifically, we say that fn converges pointwise to fwhen fn(x) → f(x) for each x ∈ X. Our argument above shows that uni-form convergence implies pointwise convergence. However, the converse neednot hold, i.e., pointwise convergence does not imply uniform convergence ingeneral. Here is an example.

Fig. 2.1 Graphs of the functions f2 (dashed) and f10 (solid) from Example 2.9.1.

Example 2.9.1 (Shrinking Triangles). Set X = [0, 1].Given n ∈ N, define

fn(x) =

0, x = 0,

linear, 0 < x < 12n ,

1, x = 12n ,

linear, 12n < x < 1

n ,

0, 1n ≤ x ≤ 1.

Each function fn is continuous on [0, 1], and fn(x) → 0 for every x ∈ [0, 1](see the illustration in Figure 2.1). However, fn does not converge uniformlyto the zero function on [0, 1] because

∀n ∈ N, ‖0 − fn‖u = supx∈[0,1]

|0 − fn(x)| = 1. ♦

According to the following lemma, the uniform limit of a sequence of con-tinuous functions is continuous.

Lemma 2.9.2. Assume that X is a metric space. Let fn ∈ Cb(X) for n ∈N be given, and let f : X → C be any scalar-valued function on X. If fn

converges uniformly to f, then f ∈ Cb(X).

Proof. Let x be any fixed point in X, and choose ε > 0. Then, by definitionof uniform convergence, there exists some integer n > 0 such that

Page 35: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.9 The Space Cb(X) 31

‖f − fn‖u < ε. (2.6)

In fact, equation (2.6) will be satisfied for all large enough n, but we needonly one particular n for this proof.

The function fn is continuous, so there is a δ > 0 such that for all y ∈ Iwe have

d(x, y) < δ =⇒ |fn(x) − fn(y)| < ε.

Consequently, if y ∈ X is any point that satisfies |x − y| < δ, then

|f(x) − f(y)| ≤ |f(x) − fn(x)| + |fn(x) − fn(y)| + |fn(y) − f(y)|

≤ ‖f − fn‖u + ε + ‖fn − f‖u

< ε + ε + ε = 3ε. (2.7)

This proves that f is continuous. To see that f is bounded, observe thatfn is bounded because it belongs to Cb(X), and f − fn is bounded because‖f − fn‖u < ε. Since the sum of two bounded functions is bounded, weconclude that f = fn + (f − fn) is bounded. Thus f is both bounded andcontinuous, so f ∈ Cb(X). ⊓⊔

Next, we use Lemma 2.9.2 to prove that Cb(X) is complete with respect tothe uniform norm.

Theorem 2.9.3. If {fn}n∈N is a sequence in Cb(X) that is Cauchy withrespect to ‖·‖u, then there exists a function f ∈ Cb(X) such that fn convergesto f uniformly. Consequently Cb(X) is a Banach space with respect to theuniform norm.

Proof. Suppose that {fn}n∈N is a uniformly Cauchy sequence in Cb(X). Ifwe fix any particular point x ∈ X, then for all m and n we have

|fm(x) − fn(x)| ≤ ‖fm − fn‖u.

It follows that {fn(x)}n∈N is a Cauchy sequence of scalars. Since R and C

are complete, this sequence of scalars must converge. Define

f(x) = limn→∞

fn(x), x ∈ X.

Now choose ε > 0. Then there exists an N such that ‖fm − fn‖u < ε forall m, n > N. If n > N, then for every x ∈ X we have

|f(x) − fn(x)| = limm→∞

|fm(x) − fn(x)| ≤ limm→∞

‖fm − fn‖u ≤ ε.

Hence ‖f − fn‖u ≤ ε whenever n > N. This shows that fn → f uniformlyas n → ∞. Consequently Lemma 2.9.2 implies that f ∈ Cb(X). Thus everyuniformly Cauchy sequence in Cb(X) converges uniformly to a function in

Page 36: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

32 2 Norms and Banach Spaces

Cb(X). Therefore we have shown that Cb(X) is complete with respect to theuniform norm. ⊓⊔

The Weierstrass Approximation Theorem is an important result aboutCb(X) when X = [a, b] is a finite closed interval in the real line. This theo-rem, which we state next, tells us that the set of polynomial functions is adense subset of C[a, b]. There are many different proofs of the WeierstrassApproximation Theorem; one can be found in [Rud76, Thm. 7.26].

Theorem 2.9.4 (Weierstrass Approximation Theorem). Given a func-tion f ∈ C[a, b] and given ε > 0, there exists some polynomial

p(x) =N

k=0

ckxk

such that‖f − p‖u = sup

x∈[a,b]

|f(x) − p(x)| < ε. ♦

We will use the Weierstrass Approximation Theorem to help illustrate thedifference between a Hamel basis, a complete sequence, and a Schauder basis.

Example 2.9.5. (a) Let

M = {xk}∞k=0 = {1, x, x2, . . . }.

By definition, a polynomial is a finite linear combination of the monomials1, x, x2, . . . , so the set of polynomials is the finite linear span of M:

P = span(M).

SinceN

k=0

ckxk = 0 ⇐⇒ c0 = · · · = cN = 0,

the set of monomials M is finitely linearly independent. Hence M is a Hamelbasis for P.

(b) Restricting our attention to functions on the domain [a, b], we havethat span(M) = P is a proper subset of C[a, b]. Therefore M is not a Hamelbasis for C[a, b]. On the other hand, by combining Theorem 2.9.4 with Lemma1.3.1 we see that span(M) is dense in C[a, b], i.e.,

span(M) = C[a, b].

Thus M is a complete sequence in C[a, b] in the sense of Definition 2.7.1.

(c) Even though M is complete and finitely linearly independent, not everyfunction f ∈ C[a, b] can be written as

Page 37: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.9 The Space Cb(X) 33

f(x) =

∞∑

k=0

αkxk. (2.8)

A series of this form is called a power series, and if it converges at some pointx, then it converges absolutely for all points t with |t| < r where r = |x|. Infact (see Problem 2.9.18), the function f so defined is infinitely differentiableon (−r, r). Therefore a continuous function that is not infinitely differentiablecannot be written as a power series (for example, consider f(x) = |x−c| wherea < c < b). Although such a function belongs to the closed span of M, itcannot be written in the form given in equation (2.8). Consequently M isnot a Schauder basis for C[a, b]. ♦

Problems

2.9.6. Prove Lemma 2.3.1.

2.9.7. Prove that every open ball Br(x) in a normed space X is convex.

2.9.8. Let Y be a subspace of a Banach space X, and let the norm on Y bethe restriction of the norm on X to the set Y. Prove that Y is a Banach spacewith respect to this norm if and only if Y is a closed subset of X.

2.9.9. Suppose that {xn}n∈N is a Cauchy sequence in a normed space X.Show that there exists a subsequence {xnk

}k∈N such that ‖xnk+1−xnk

‖ < 2−k

for every k.

2.9.10. Suppose that A = {xn}n∈N is a Schauder basis for a Banach space X.Prove that equality holds in equation (2.3).

2.9.11. Suppose that {xn}n∈N is a sequence of vectors in ℓ1. For each n, writethe components of xn as

xn =(

xn(k))

k∈N=

(

xn(1), xn(2), . . .)

.

Suppose thatx =

(

x(k))

k∈N=

(

x(1), x(2), . . .)

is such that xn converges to x in ℓ1-norm, i.e., ‖x − xn‖1 → 0 as n → ∞.Prove that for each fixed k, the kth component of xn converges to the kthcomponent of x, i.e.,

limk→∞

xn(k) = x(k), k ∈ N.

Thus, convergence in ℓ1-norm implies componentwise convergence.

Page 38: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

34 2 Norms and Banach Spaces

2.9.12. The closed unit ball in the space ℓ1 is D ={

x ∈ ℓ1 : ‖x‖1 ≤ 1}

.

(a) Prove that D is a closed and bounded subset of ℓ1.

(b) Let δn be the nth standard basis vector (see Notation 1.1.3). Provethat the set of standard basis vectors {δn}n∈N contains no convergent sub-sequences. Conclude that D is not sequentially compact, and therefore (byTheorem 1.4.5) D is not compact.

2.9.13. Let S be the set of all “finite sequences” whose components are ra-tional, i.e.,

S ={

(r1, . . . , rN , 0, 0, . . . ) : N > 0, r1, . . . , rN rational}

,

where we say that a complex number is rational if both its real and imaginaryparts are rational. Prove that S is a countable, dense subset of ℓ1. Concludethat ℓ1 is separable.

2.9.14. Suppose that X is a Banach space that has a Schauder basis {xn}n∈N.Prove that X must be separable.

2.9.15. For the following choices of functions gn, determine whether the se-quence {gn}n∈N is: pointwise convergent, uniformly convergent, uniformlyCauchy, or bounded with respect to the uniform norm.

(a) gn(x) = nfn(x), x ∈ [0, 1], where fn is the Shrinking Triangle fromExample 2.9.1.

(b) gn(x) = e−n|x|, x ∈ R.

(c) gn(x) = xe−n|x|, x ∈ R.

(d) gn(x) =nx

1 + n2x2, x ∈ R.

2.9.16. The space of continuous functions that “vanish at infinity” is

C0(R) ={

f ∈ C(R) : limx→±∞

f(x) = 0}

.

Prove the following statements.

(a) C0(R) is a closed subspace of Cb(R) with respect to the uniform norm,and is therefore a Banach space with respect to ‖ · ‖u.

(b) Every function in C0(R) is uniformly continuous on R, but there existfunctions in Cb(R) that are not uniformly continuous.

(c) If f ∈ C0(R) and g(x) = f(x − k) where k is fixed, then f ∈ C0(R)and ‖g‖u = ‖f‖u.

(d) The “closed unit ball” D = {f ∈ C0(R) : ‖f‖u ≤ 1} in C0(R) is closedand bounded, but it is not sequentially compact.

Page 39: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

2.9 The Space Cb(X) 35

2.9.17. Let

fn(x) =e−n2x2

n2, x ∈ R.

(a) Prove that the series∑∞

n=1 fn converges absolutely in Cb(R) withrespect to the uniform norm, and therefore converges since Cb(R) is a Banachspace.

(b) Prove that the series∑∞

n=1 f ′n converges in Cb(R) with respect to the

uniform norm, but it does not converge absolutely with respect to that norm.

2.9.18. Let (ck)k≥0 be a fixed sequence of scalars. Suppose that the series∑∞

k=0 ckxk converges for some number x ∈ R, and set r = |x|. Prove thefollowing statements.

(a) The series f(t) =∑∞

k=0 cktk converges absolutely for all t ∈ (−r, r),i.e.,

∞∑

k=0

|ck| |t|k < ∞, |t| < r.

(b) f is infinitely differentiable on the interval (−r, r).

Page 40: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric
Page 41: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Chapter 3

Inner Products and Hilbert Spaces

In a normed vector space, each vector has an assigned length, and fromthis we obtain the distance from x to y as the length of the vector x − y.For vectors in Rd and Cd we also know how to measure the angle betweenvectors. In particular, two vectors x, y in Rd or Cd are perpendicular, ororthogonal, if their dot product is zero. An inner product is a generalizationof the dot product, and it provides us with a generalization of the notion oforthogonality to vector spaces other than Rd or Cd.

3.1 The Definition of an Inner Product

Here are the defining properties of an inner product.

Definition 3.1.1 (Semi-Inner Product, Inner Product). Let H be avector space over either the real field R or the complex field C. A semi-innerproduct on H is a scalar-valued function 〈·, ·〉 on H × H such that for allvectors x, y, z ∈ H and all scalars a, b we have:

(a) Nonnegativity: 〈x, x〉 ≥ 0,

(b) Conjugate Symmetry: 〈x, y〉 = 〈y, x〉, and

(c) Linearity in the First Variable: 〈ax + by, z〉 = a〈x, z〉 + b〈y, z〉.

If a semi-inner product 〈·, ·〉 also satisfies:

(d) Uniqueness: 〈x, x〉 = 0 if and only if x = 0,

then it is called an inner product on H. In this case, H is called an innerproduct space or a pre-Hilbert space. ♦

The usual dot product

u · v = u1v1 + · · · + unvn (3.1)

c©2014 Christopher Heil37

Page 42: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

38 3 Inner Products and Hilbert Spaces

is an inner product on Rd (using real scalars) and on Cn (using complexscalars). Of course, if the scalar field is R, then the complex conjugate inequation (3.1) is superfluous. Similarly, if H is a real vector space then thecomplex conjugate in the definition of conjugate symmetry is irrelevant since〈x, y〉 will be real.

If 〈·, ·〉 is a semi-inner product on a vector space H, then we define

‖x‖ = 〈x, x〉1/2, x ∈ H.

We will prove in Lemma 3.2.3 that ‖ · ‖ is a seminorm on H, and thereforewe refer to ‖ · ‖ as the seminorm induced by 〈·, ·〉. Likewise, if 〈·, ·〉 is an innerproduct, then we will prove that ‖ · ‖ is a norm on H, and in this case werefer to ‖ · ‖ as the norm induced by 〈·, ·〉. It may be possible to place othernorms on H, but unless we explicitly state otherwise, we assume that allnorm-related statements on an inner product space are taken with respect tothe induced norm.

3.2 Properties of an Inner Product

Here are some useful properties of inner products (the proof is assigned asProblem 3.9.4).

Lemma 3.2.1. If 〈·, ·〉 is a semi-inner product on a vector space H, then thefollowing statements hold for all vectors x, y, z ∈ H and all scalars a, b.

(a) Antilinearity in the Second Variable: 〈x, ay + bz〉 = a 〈x, y〉 + b 〈x, z〉.

(b) Polar Identity: ‖x + y‖2 = ‖x‖2 + 2Re(〈x, y〉) + ‖y‖2.

(c) Pythagorean Theorem: If 〈x, y〉 = 0, then ‖x ± y‖2 = ‖x‖2 + ‖y‖2.

(d) Parallelogram Law: ‖x + y‖2 + ‖x − y‖2 = 2(

‖x‖2 + ‖y‖2)

. ♦

The inequality that we prove next is variously known as the Schwarz,Cauchy–Schwarz, Cauchy–Bunyakovski–Schwarz, or CBS Inequality.

Theorem 3.2.2 (Cauchy–Bunyakovski–Schwarz Inequality). If 〈·, ·〉 isa semi-inner product on a vector space H, then

∀x, y ∈ H, |〈x, y〉| ≤ ‖x‖ ‖y‖.

Proof. If x = 0 or y = 0 then there is nothing to prove, so suppose that xand y are both nonzero. Let α ∈ C be a scalar such that |α| = 1 and

〈x, y〉 = α |〈x, y〉|.

Then for each t ∈ R, the Polar Identity implies that

Page 43: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.3 Hilbert Spaces 39

0 ≤ ‖x − αty‖2 = ‖x‖2 − 2Re(〈x, αty〉) + t2 ‖y‖2

= ‖x‖2 − 2t Re(α 〈x, y〉) + t2 ‖y‖2

= ‖x‖2 − 2t |〈x, y〉| + t2 ‖y‖2.

This is a real-valued quadratic polynomial in the variable t. Since this poly-nomial is nonnegative, it can have at most one real root. This requires thatthe discriminant of the polynomial be at most zero, so we must have

(

−2 |〈x, y〉|)2

− 4 ‖x‖2 ‖y‖2 ≤ 0.

The desired result follows by rearranging this inequality. ⊓⊔

By combining the Polar Identity with the Cauchy–Bunyakovski–SchwarzInequality, we can now prove that the induced seminorm ‖ · ‖ satisfies theTriangle Inequality.

Lemma 3.2.3. Let H be a vector space. If 〈·, ·〉 is a semi-inner product onH, then ‖ · ‖ is a seminorm on H, and if 〈·, ·〉 is an inner product on H, then‖ · ‖ is a norm on H.

Proof. The only property of a seminorm that is not obvious is the Trian-gle Inequality. To prove this, we use the Polar Identity and the Cauchy–Bunyakovski–Schwarz Inequality to compute that

‖x + y‖2 = ‖x‖2 + 2Re〈x, y〉 + ‖y‖2

≤ ‖x‖2 + 2 |〈x, y〉| + ‖y‖2

≤ ‖x‖2 + 2 ‖x‖ ‖y‖ + ‖y‖2

=(

‖x‖ + ‖y‖)2

.

The Triangle Inequality follows by taking square roots. ⊓⊔

3.3 Hilbert Spaces

The question of whether every Cauchy sequence in a given inner productspace must converge is very important, just as it is in a metric or normedspace. We give the following name to those inner product spaces that havethis property.

Definition 3.3.1 (Hilbert Space). An inner product space H is called aHilbert space if it is complete with respect to the induced norm. ♦

Thus, an inner product space is a Hilbert space if and only if every Cauchysequence in H converges to an element of H. Equivalently, a Hilbert space is

Page 44: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

40 3 Inner Products and Hilbert Spaces

an inner product space that is a Banach space with respect to the inducednorm.

Using real scalars, Rd is a Hilbert space with respect to the usual dotproduct. Likewise, Cd is a Hilbert space with respect to the dot product ifwe use complex scalars.

Here is an example of an infinite-dimensional Hilbert space (compare thisto Example 1.1.2, which introduced the space ℓ1).

Example 3.3.2. Given a sequence of real or complex scalars

x = (xk)k∈N = (x1, x2, . . . ),

we define the ℓ2-norm of x to be

‖x‖2 = ‖(xk)k∈N‖2 =

( ∞∑

k=1

|xk|2

)1/2

. (3.2)

Note that we have not yet proved that ‖ · ‖2 is a norm in any sense; we willaddress this issue below.

We say that a sequence x = (xk)k∈N is square summable if ‖x‖2 < ∞, andwe let ℓ2 denote the space of all square summable sequences, i.e.,

ℓ2 =

{

x = (xk)k∈N : ‖x‖22 =

∞∑

k=1

|xk|2 < ∞

}

.

We will define an inner product on ℓ2. First we recall that the arithmetic-geometric mean inequality implies that if a, b ≥ 0, then

ab ≤a2

2+

b2

2.

Consequently, if we choose any two vectors x = (xk)k∈N and y = (yk)k∈N inℓ2, then

∞∑

k=1

|xkyk| ≤

∞∑

k=1

(

|xk|2

2+

|yk|2

2

)

=‖x‖2

2

2+

‖y‖22

2< ∞. (3.3)

Therefore we can define

〈x, y〉 =

∞∑

k=1

xkyk, x, y ∈ ℓ2, (3.4)

because equation (3.3) tells us that this series of scalars converges absolutely.In particular, for x = y we have

Page 45: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.3 Hilbert Spaces 41

〈x, x〉 =

∞∑

k=1

|xk|2 = ‖x‖2

2. (3.5)

We can check that the function 〈·, ·〉 satisfies all of the requirements of aninner product. For example, simply by definition we have 0 ≤ 〈x, x〉 < ∞for every x ∈ ℓ2. Further, if 〈x, x〉 = 0 then it follows from equation (3.5)that xk = 0 for every k, so x = 0. This establishes the nonnegativity anduniqueness requirements of an inner product, and the conjugate symmetryand linearity in the first variable requirements are easily checked as well.

Thus 〈·, ·〉 is an inner product on ℓ2, and therefore ℓ2 is an inner productspace with respect to this inner product. Further, equation (3.5) shows thatthe norm induced from this inner product is precisely the ℓ2-norm ‖ · ‖2, soLemma 3.2.3 implies that ‖ · ‖2 really is a norm on ℓ2.

Is ℓ2 a Hilbert space? This is slightly more difficult to check, but an argu-ment very similar to the one used in Theorem 1.2.6 shows that every Cauchysequence in ℓ2 converges to an element of ℓ2 (we assign the details as Problem3.9.9). Consequently ℓ2 is complete, so it is a Hilbert space. ♦

The space ℓ1 is a proper subspace of ℓ2. For, if x = (xk)k∈N belongs to ℓ1

then we must have |xk| < 1 for all large enough k, say k > N, and therefore

∞∑

k=N+1

|xk|2 ≤

∞∑

k=N+1

|xk| ≤ ‖x‖1 < ∞.

Hence x ∈ ℓ2. On the other hand, the sequence

x =(

1k

)

k∈N=

(

1, 12 , 1

3 , . . .)

belongs to ℓ2, but it does not belong to ℓ1. Now, since ℓ1 is a subspace ofℓ2, we can define an inner product on ℓ1 simply by restricting equation (3.4)to vectors in ℓ1. However, the norm induced from this inner product is ‖ · ‖2

rather than ‖·‖1. Although ℓ1 is complete with respect to the norm ‖·‖1, it isnot complete with respect to the norm ‖ · ‖2 (see Problem 3.9.10). Thereforeℓ1 is an inner product space with respect to the inner product defined inequation (3.4), but it is not a Hilbert space.

Here is a different example of an inner product space.

Example 3.3.3. Let C[a, b] be the space of all continuous functions of the formf : [a, b] → C. All functions in C[a, b] are bounded and Riemann integrable.If we choose f, g ∈ C[a, b], then the product f(x) g(x) is continuous andRiemann integrable, so we can define

〈f, g〉 =

∫ b

a

f(x) g(x) dx, f, g ∈ C[a, b]. (3.6)

Page 46: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

42 3 Inner Products and Hilbert Spaces

We can easily see that 〈·, ·〉 satisfies the nonnegativity, conjugate symmetry,and linearity in the first variable requirements stated in Definition 3.1.1, andhence is at least a semi-inner product on C[a, b]. According to Problem 3.9.8,the uniqueness requirement also holds, and therefore 〈·, ·〉 is an inner producton C[a, b]. The norm induced from this inner product is

‖f‖2 = 〈f, f〉1/2 =

(∫ b

a

|f(x)|2 dx

)1/2

, f ∈ C[a, b].

We call ‖f‖2 the L2-norm of the function f. ♦

Although we have defined an inner product on C[a, b], we will prove thatC[a, b] is not complete with respect to the norm ‖ · ‖2 that is induced fromthat inner product (in contrast, note that Theorem 2.9.3 implies that C[a, b]is complete with respect to a different norm—the uniform norm ‖ · ‖u).

Lemma 3.3.4. C[a, b] is not a Hilbert space with respect to 〈·, ·〉.

Proof. For simplicity of presentation we will take a = −1 and b = 1. Asillustrated in Figure 3.1 for the case n = 5, define

fn(x) =

−1, −1 ≤ x ≤ − 1n ,

linear, − 1n < x < 1

n ,

1, 1n ≤ x ≤ 1;

Fig. 3.1 Graph of the function f5.

Given positive integers m < n we compute that

‖fm − fn‖22 =

∫ 1

−1

|fm(x) − fn(x)|2 dx

=

∫ 1/m

−1/m

|g(x) − fn(x)|2 dx

∫ 1/m

−1/m

1 dx =2

m.

Page 47: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.3 Hilbert Spaces 43

Therefore, if we fix ε > 0 then for all m,n large enough we will have‖fm − fn‖2 < ε. This shows that {fn}n∈N is a Cauchy sequence in C[−1, 1].However, we will prove that there is no function g ∈ C[−1, 1] such that‖g − fn‖2 → 0 as n → ∞.

Suppose that there was some g ∈ C[−1, 1] such that ‖g − fn‖2 → 0. Fix0 < c < 1, and suppose that h is not identically 1 on [c, 1]. Then

C =

∫ 1

c

|g(x) − 1| dx > 0.

On the other hand, fn is identically 1 on [c, 1] for all n > 1/c, so for all largeenough n we have

‖g − fn‖22 =

∫ 1

−1

|g(x) − fn(x)|2 dx

∫ 1

c

|g(x) − fn(x)|2 dx

=

∫ 1

c

|g(x) − 1| dx ≥ C.

Since C > 0, it follows that ‖g−fn‖2 →/ 0, which is a contradiction. Thereforeg must be identically 1 on [c, 1]. This is true for every c > 1, so g(x) = 1 forall 0 < x ≤ 1. A similar argument shows that g(x) = −1 for all −1 ≤ x < 0.There is no continuous function that takes these values, so we have obtaineda contradiction.

Thus, although {fn}n∈N is Cauchy in C[−1, 1], there is no function inC[−1, 1] that this sequence can converge to with respect to the induced norm‖ · ‖2. Therefore C[−1, 1] is not complete with respect to this norm. ⊓⊔

Each incomplete normed space that we encountered prior to C[a, b] wascontained in some larger complete space. For example, c00 is incomplete withrespect to the norm ‖ · ‖1, but it is a subset of ℓ1, which is complete withrespect to ‖ · ‖1. Likewise, Cc(R) is incomplete with respect to the uniformnorm, yet it is contained in C0(R), which is a Banach space with respectto that norm. It is likewise true that the inner product space C[a, b] is con-tained in a larger Hilbert space, the Lebesgue space L2[a, b]. However, definingL2[a, b] is beyond the scope of this manuscript, as it requires the theory ofthe Lebesgue integral. We refer to texts such as [Fol99], [SS05], or [WZ77] fordiscussion of the Lebesgue integral and the Hilbert space L2[a, b].

Page 48: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

44 3 Inner Products and Hilbert Spaces

3.4 Orthogonal and Orthonormal Sets

Orthogonality plays a important role in the analysis of inner product spaces.We declare that two vectors are orthogonal if their inner product is zero (con-sequently, the zero vector is orthogonal to every other vector). We say that acollection of vectors is orthogonal if every pair of vectors from the collectionis orthogonal. Often it is convenient to work with orthogonal unit vectors;we refer to these as orthonormal vectors. Here are the precise definitions (the“Kronecker delta” is defined by δij = 1 if i = j, and δij = 0 if i 6= j).

Definition 3.4.1. Let H be an inner product space, and let I be an arbitraryindex set.

(a) Two vectors x, y ∈ H are orthogonal, denoted x ⊥ y, if 〈x, y〉 = 0.

(b) A set of vectors {xi}i∈I is orthogonal if 〈xi, xj〉 = 0 whenever i 6= j.

(c) A set of vectors {xi}i∈I is orthonormal if it is orthogonal and each vectorxi is a unit vector. Using the Kronecker delta notation, {xi}i∈I is anorthonormal set if for all i, j ∈ I we have

〈xi, xj〉 = δij =

{

1, i = j,

0, i 6= j.♦

The zero vector may be an element of a set of orthogonal vectors. However,any orthogonal set {xi}i∈N of nonzero vectors can be rescaled to form anorthonormal set. That is, if we set

yi =xi

‖xi‖,

then {yi}i∈I is an orthonormal set.

Example 3.4.2. The sequence of standard basis vectors {δn}n∈N is an ortho-normal sequence in ℓ2. ♦

3.5 Orthogonal Complements

We define the orthogonal complement of a set A to be the largest set B thatis orthogonal to A.

Definition 3.5.1 (Orthogonal Complement). Let A be a subset of aninner product space H. The orthogonal complement of A is

A⊥ ={

x ∈ H : x ⊥ A}

={

x ∈ H : 〈x, y〉 = 0 for all y ∈ A}

. ♦

Page 49: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.6 Orthogonal Projections 45

Here are some properties of orthogonal complements (the proof is assignedas Problem 3.9.12).

Lemma 3.5.2. If A is a subset of an inner product space H, then the follow-ing statements hold.

(a) A⊥ is a closed subspace of H.

(b) H⊥ = {0} and {0}⊥ = H.

(c) If A ⊆ B, then B⊥ ⊆ A⊥.

(d) A ⊆ (A⊥)⊥. ♦

Later we will prove that if M is a closed subspace of a Hilbert space, then(M⊥)⊥ = M (see Lemma 3.6.4).

3.6 Orthogonal Projections

The following theorem states that if S is a closed and convex subset of aHilbert space H, then for each vector x ∈ H there exists is a unique vectory ∈ S that is closest to x. Looking at the proof, we can see why we mustassume in this result that H is a Hilbert space and not just an inner productspace—we need to know that every Cauchy sequence in H will converge toan element of H.

Theorem 3.6.1 (Closest Point Theorem). Let H be a Hilbert space, andlet S be a nonempty closed, convex subset of H. Given any x ∈ H there existsa unique vector y ∈ S that is closest to x. That is, there is a unique vectory ∈ S that satisfies

‖x − y‖ = dist(x, S) = inf{

‖x − k‖ : k ∈ S}

.

Proof. Set d = dist(x, S). Then, by definition of infimum, there exist vectorsyn ∈ S such that ‖x − yn‖ → d as n → ∞, and for each of these vectors wehave ‖x − yn‖ ≥ d. Therefore, if we fix ε > 0 then we can find an integerN > 0 such that

d2 ≤ ‖h − yn‖2 ≤ d2 + ε2 for all n > N.

Set

p =ym + yn

2.

Since p is the midpoint of the line segment joining ym to yn we have p ∈ S,and therefore

‖x − p‖ ≥ dist(x, S) = d.

Using the Parallelogram Law, it follows that if m, n > N, then

Page 50: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

46 3 Inner Products and Hilbert Spaces

‖yn − ym‖2 + 4d2 ≤ ‖yn − ym‖2 + 4 ‖x − p‖2

= ‖(x − yn) − (x − ym)‖2 + ‖(x − yn) + (x − ym)‖2

= 2(

‖x − yn‖2 + ‖x − ym‖2

)

≤ 4 (d2 + ε2).

Rearranging, we see that ‖ym−yn‖ ≤ 2ε for all m, n > N. Therefore {yn}n∈N

is a Cauchy sequence in H. Since H is complete, this sequence must converge,say to y. Since S is closed and yn ∈ S for every n, the vector y must belongto S. Also, since x − yn → x − y, it follows from the continuity of the normthat

‖x − y‖ = limn→∞

‖x − yn‖ = d.

Hence y is a point in S that is closest to x.It only remains to show that y is the unique point in S that is closest

to x. If z ∈ S is also a closest point, then ‖x − y‖ = d = ‖x − z‖. Further,the midpoint p = (y + z)/2 belongs to S, so ‖x − p‖ ≥ d. Applying theParallelogram Law again, we see that

4d2 = 2(

‖x − y‖2 + ‖x − z‖2)

= ‖(x − y) − (x − z)‖2 + ‖(x − y) + (x − z)‖2

= ‖y − z‖2 + 4 ‖x − p‖2

≥ ‖y − z‖2 + 4d2.

Rearranging this yields ‖y − z‖ ≤ 0, which implies that y = z. ⊓⊔

In particular, every closed subspace M of H is nonempty, closed, andconvex. For this setting we introduce a name for the point in M that is closestto a given vector x. We also use the same name to denote the function thatmaps x to the point in M that is closest to x.

Definition 3.6.2 (Orthogonal Projection). Let M be a closed subspaceof a Hilbert space H.

(a) Given x ∈ H, the unique vector p ∈ M that is closest to x is called theorthogonal projection of x onto M.

(b) The function P : H → H defined by Px = p, where p is the orthog-onal projection of x onto M, is called the orthogonal projection of Honto M. ♦

Since the orthogonal projection p is the vector in M that is closest to x, wecan think of p as being the best approximation to x by vectors from M. Thedifference vector e = x − p is the error in this approximation. The followinglemma states that the orthogonal projection of x is the unique vector p ∈ Hsuch that the error vector e = x − p is orthogonal to M.

Page 51: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.6 Orthogonal Projections 47

Lemma 3.6.3. Let M be a closed subspace of a Hilbert space H. Given vec-tors x and p in H, the following four statements are equivalent.

(a) p is the orthogonal projection of x onto M, i.e., p is the unique point inM that is closest to h.

(b) p ∈ M and x − p ⊥ M.

(c) x = p + e where p ∈ M and e ∈ M⊥.

(d) e = x − p is the orthogonal projection of x onto M⊥.

Proof. We will prove one implication, and assign the task of proving theremaining (easier) implications as Problem 3.9.13. We assume that scalarsin this problem are complex, but the proof remains valid if we assume thatscalars are real.

(a) ⇒ (b). Let p be the (unique) point in M closest to x, and let e = p−x.Choose any vector y ∈ M. We must show that 〈y, e〉 = 0. Since M is asubspace, p + λy ∈ M for any scalar λ ∈ C. Hence,

‖x − p‖2 ≤ ‖x − (p + λy)‖2 = ‖(x − p) − λy‖2

= ‖x − p‖2 − 2Re〈λy, x − p〉 + |λ|2 ‖y‖2

= ‖x − p‖2 − 2Re(λ〈y, e〉) + |λ|2 ‖y‖2.

Therefore,∀λ ∈ C, 2Re(λ〈y, e〉) ≤ |λ|2 ‖y‖2.

If we consider λ = t > 0, then we can divide through by t to get

∀ t > 0, 2Re〈y, e〉 ≤ t ‖y‖2.

Letting t → 0+, we conclude that Re〈y, e〉 ≤ 0. If we similarly take λ = t < 0and let t → 0−, we obtain Re〈y, e〉 ≥ 0, so Re〈y, e〉 = 0.

Finally, by taking λ = it with t > 0 and then λ = it with t < 0, and thenλ = it with t < 0, it follows that Im(〈y, e〉) = 0. ⊓⊔

We will use Lemma 3.6.3 to compute the orthogonal complement of theorthogonal complement of a set.

Lemma 3.6.4. Let H be a Hilbert space.

(a) If M is a closed subspace of H, then (M⊥)⊥ = M.

(b) If A is any subset of H, then

A⊥ = span(A)⊥ = span(A)⊥ and (A⊥)⊥ = span(A).

Proof. (a) We are given a closed subspace M in H. If f ∈ M then 〈f, g〉 = 0for every g ∈ M⊥, so f ∈ (M⊥)⊥. Hence M ⊆ (M⊥)⊥.

Conversely, suppose that f ∈ (M⊥)⊥. Let p be the orthogonal projectionof f onto M. Since M is a closed subspace, we have f = p + e where p ∈ M

Page 52: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

48 3 Inner Products and Hilbert Spaces

and e ∈ M⊥. Since f and p belong to M and we have seen that M ⊆ (M⊥)⊥,it follows that e = f − p ∈ (M⊥)⊥. However, we also know that e ∈ M⊥,so e is orthogonal to itself and therefore is zero. Hence f = p + 0 ∈ M. Thisshows that (M⊥)⊥ ⊆ M.

(b) Now we are given an arbitrary subset A of H. Let M = span(A). Wemust show that A⊥ = M⊥. Since A ⊆ M, we have M⊥ ⊆ A⊥.

Suppose that f ∈ A⊥. Then f ⊥ A, i.e., f is orthogonal to every vectorin A. By forming linear combinations, it follows that f ⊥ span(A). By takinglimits, it follows from this that f ⊥ span(A) = M. Hence f ∈ M⊥ andtherefore A⊥ ⊆ M⊥.

Thus, we have shown that A⊥ = M⊥. Applying part (a), we conclude that(A⊥)⊥ = (M⊥)⊥ = M. ⊓⊔

By taking A to be a sequence, we obtain the following useful corollary.

Corollary 3.6.5. Given a sequence {xn}n∈N in a Hilbert space H, the fol-lowing two statements are equivalent.

(a) {xn}n∈N is a complete sequence, i.e., span{xn}n∈N is dense in H.

(b) The only vector in H that is orthogonal to every xn is the zero vector,i.e.,

x ∈ H and 〈x, xn〉 = 0 for every n =⇒ x = 0.

Proof. By definition, {xn} is complete if and only if its finite linear span is

dense in H. Lemma 3.6.4 tells us that span{xn} = {xn}⊥, so we conclude

that {xn} is complete if only if {xn}⊥ = {0}. ⊓⊔

3.7 Orthonormal Sequences

If {en}n∈N is any sequence of orthonormal vectors in a Hilbert space, thenits closed span M = span

(

{en}n∈N

)

is a closed subspace of H. The followinglemma gives an explicit formula for the orthogonal projection of a vectoronto M, along with other useful properties of {en}n∈N Entirely similar resultshold for a finite orthonormal sequence {e1, . . . , eN} by replacing n = 1, 2, . . .by n = 1, . . . , N in Theorem 3.7.1. In fact, the proof is easier for finitesequences, since in that case there are no issues about convergence of infiniteseries.

Theorem 3.7.1. Let H be a Hilbert space, let {en}n∈N be an orthonormalsequence in H, and set

M = span(

{en}n∈N

)

. (3.7)

Then the following statements hold.

Page 53: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.7 Orthonormal Sequences 49

(a) Bessel’s Inequality:

∞∑

n=1

|〈x, en〉|2 ≤ ‖x‖2 for all x ∈ H.

(b) If cn is a scalar for n ∈ N and the series x =∑∞

n=1 cnen converges, then

cn = 〈x, en〉, all n ∈ N.

(c) If cn is a scalar for n ∈ N, then

∞∑

n=1

cnen converges ⇐⇒

∞∑

n=1

|cn|2 < ∞.

Further, in this case the series∑∞

n=1 cnen converges unconditionally, i.e.,it converges regardless of the ordering of the index set.

(d) If x ∈ H, then

p =

∞∑

n=1

〈x, en〉 en

is the orthogonal projection of x onto M, and

‖p‖2 =

∞∑

n=1

|〈x, en〉|2.

(e) If x ∈ H, then

x ∈ M ⇐⇒ x =∞∑

n=1

〈x, en〉 en ⇐⇒ ‖x‖2 =∞∑

n=1

|〈x, en〉|2. (3.8)

Proof. (a) Choose x ∈ H. For each N ∈ N define

yN = x −

N∑

n=1

〈x, xn〉xn.

If 1 ≤ m ≤ N, then

〈yN , xm〉 = 〈x, xm〉 −N

n=1

〈x, xn〉 〈xn, xm〉 = 〈x, xm〉 − 〈x, xm〉 = 0.

Thus {x1, . . . , xN , yN} is an orthogonal sequence in H (though not necessar-ily orthonormal, since we do not know the norm of yN ). Therefore, by thePythagorean Theorem (see Problem 3.9.7),

Page 54: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

50 3 Inner Products and Hilbert Spaces

‖x‖2 =∥

∥yN +

N∑

n=1

〈x, xn〉xn

2

= ‖yN‖2 +

N∑

n=1

‖〈x, xn〉xn‖2

= ‖yN‖2 +

N∑

n=1

|〈x, xn〉|2

N∑

n=1

|〈x, xn〉|2.

Letting N → ∞, we obtain Bessel’s Inequality.

(b) If x =∑

cnxn converges, then for each fixed m we have by continuityof the inner product that

〈x, xm〉 =∞∑

n=1

cn 〈xn, xm〉 =∞∑

n=1

cnδmn = cm.

(c) If∑

cnxn converges, then part (b) implies that cn = 〈x, xn〉, and hence∑

|cn|2 < ∞ by Bessel’s Inequality.

For the converse direction, suppose that∑

|cn|2 < ∞. Set

sN =

N∑

n=1

cnxn and tN =

N∑

n=1

|cn|2.

We know that {tN}N∈N is a convergent (hence Cauchy) sequence of scalars,and we must show that {sN}N∈N is a convergent sequence of vectors. Wehave for N > M that

‖sN − sM‖2 =∥

N∑

n=M+1

cnxn

2

=

N∑

n=M+1

‖cnxn‖2

=N

n=M+1

|cn|2

= |tN − tM |.

Since {tN}N∈N is Cauchy, we conclude that {sN}N∈N is a Cauchy sequencein H and hence converges.

(d) By Bessel’s Inequality and part (c), we know that the series

Page 55: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.7 Orthonormal Sequences 51

p =

∞∑

n=1

〈x, xn〉xn

converges. Our task is to show that this vector p is the orthogonal projectionof x onto M.

Given any fixed k, we compute that

〈x − p, xk〉 = 〈x, xk〉 −

⟨ ∞∑

n=1

〈x, xn〉xn, xk

= 〈x, xk〉 −∞∑

n=1

〈x, xn〉 〈xn, xk〉

= 〈x, xk〉 − 〈x, xk〉 = 0,

where we have used Problem 3.9.6 to move the infinite series to the outsideof the inner product. Thus x − p is orthogonal to each vector xk. Since theinner product is antilinear in the second variable, this implies that x − p isorthogonal to every finite linear combination of the xk. Taking limits andapplying the continuity of the inner product again, it follows that x − p isorthogonal to every vector in M, i.e., x − p ⊥ M. Lemma 3.6.3 thereforeimplies that that p is the orthogonal projection of x onto M.

(e) Statement (d) tells us that p =∑

〈x, en〉 en is the orthogonal projectionof x onto M, and that

‖p‖2 = 〈p, p〉 =

∞∑

n=1

|〈x, en〉|2.

Let i, ii, iii denote the three statements that appear in equation (3.8). Wemust prove that these statements i, ii, and iii are equivalent.

i ⇒ ii. If x ∈ M, then the orthogonal projection of x onto M is x itself.Thus x = p =

〈x, xn〉xn.

ii ⇒ iii. If x = p then ‖x‖2 = ‖p‖2 =∑

|〈x, en〉|2.

iii ⇒ i. Suppose ‖x‖2 =∑

|〈x, en〉|2. Then since x−p ⊥ p, we can use the

Pythagorean Theorem to compute that

‖x‖2 = ‖(x − p) + p‖2 = ‖x − p‖2 + ‖p‖2

= ‖x − p‖2 +∞∑

n=1

|〈x, en〉|2

= ‖x − p‖2 + ‖x‖2.

Hence ‖x − p‖ = 0, so x = p ∈ M. ⊓⊔

Page 56: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

52 3 Inner Products and Hilbert Spaces

3.8 Orthonormal Bases

If the closed span of an orthonormal sequence {en}n∈N is M = H, then we saythat the sequence {en}n∈N is complete, total, or fundamental (see Definition2.7.1). Part (e) of Theorem 3.7.1 implies that if {en}n∈N is both orthonormaland complete, then every vector x ∈ H can be written as x =

〈x, en〉 en.The following theorem states that this property characterizes completeness(assuming that our sequence {en}n∈N is orthonormal), and gives several otheruseful characterizations of complete orthonormal sequences. A completelyanalogous theorem holds for finite orthonormal sequence {e1, . . . , ed} (seeProblem 3.9.15).

Theorem 3.8.1. If H is a Hilbert space and {en}n∈N is an orthonormalsequence in H, then the following five statements are equivalent.

(a) {en}n∈N is complete, i.e., span{en}n∈N = H.

(b) {en}n∈N is a Schauder basis for H, i.e., for each x ∈ H there exists aunique sequence of scalars (cn)n∈N such that x =

cnen.

(c) For each x ∈ H we have

x =

∞∑

n=1

〈x, en〉 en. (3.9)

(d) Plancherel’s Equality:

‖x‖2 =∞∑

n=1

|〈x, en〉|2 for all x ∈ H.

(e) Parseval’s Equality:

〈x, y〉 =

∞∑

n=1

〈x, en〉 〈en, y〉 for all x, y ∈ H.

Proof. (a) ⇒ (b). If {en} is complete, then its closed span is all of H, soTheorem 3.7.1(c) implies that x =

〈x, en〉 en for every x ∈ H.

(b) ⇒ (c). If statement (b) holds, then we must have cn = 〈x, xn〉 byTheorem 3.7.1(b).

(c) ⇒ (e). Suppose that statement (c) holds, and fix any vectors x, y ∈ H.Then

〈x, y〉 =⟨

∞∑

n=1

〈x, en〉 en, y⟩

=

∞∑

n=1

〈x, en〉 en, y⟩

=

∞∑

n=1

〈x, en〉 〈en, y〉,

where we have used Problem 3.9.6 to move the infinite series outside of theinner product.

Page 57: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.8 Orthonormal Bases 53

(e) ⇒ (d). This follows by taking x = y.

(d) ⇒ (c). Suppose that ‖x‖2 =∑∞

n=1 |〈x, en〉|2 for every x ∈ H. Fix x,

and define sN =∑N

n=1 〈x, en〉 en. Then, by direct calculation,

‖x − sN‖2 = ‖x‖2 − 〈x, sN 〉 − 〈sN , x〉 + ‖sN‖2

= ‖x‖2 −

N∑

n=1

|〈x, en〉|2 −

N∑

n=1

|〈x, en〉|2 +

N∑

n=1

|〈x, en〉|2

= ‖x‖2 −

N∑

n=1

|〈x, en〉|2

→ 0 as N → ∞.

Hence x =∑∞

n=1 〈x, en〉 en.

(d) ⇒ (a). Suppose that Plancherel’s Equality holds, and 〈x, en〉 = 0 forevery n ∈ N. Then

‖x‖2 =

∞∑

n=1

|〈x, en〉|2 = 0,

so x = 0. Hence {en} is complete. ⊓⊔

We will refer to a sequence that satisfies the equivalent conditions in The-orem 3.8.1 as an orthonormal basis.

Definition 3.8.2 (Orthonormal Basis). Let H be a Hilbert space. Acountably infinite orthonormal sequence {en}n∈N that is complete in H iscalled an orthonormal basis for H. ♦

We use similar terminology for finite-dimensional Hilbert spaces. Specifi-cally, if {e1, . . . , ed} is a complete orthonormal sequence in a Hilbert space H,then {e1, . . . , ed} is a basis for H in the usual vector space sense (i.e., it is aHamel basis), and we call it an orthonormal basis for H.

Example 3.8.3. The sequence of standard basis vectors {δn}n∈N is both com-plete and orthonormal in ℓ2, so it is an orthonormal basis for ℓ2. ♦

By Theorem 3.8.1, a countable sequence that is both complete and ortho-normal is a Schauder basis for H. However, we emphasize that a completesequence that is not orthonormal need not be a Schauder basis, even if it isfinitely linearly independent. An example in the Hilbert space ℓ2 is givenin Problem 3.9.21. A similar phenomenon in the Banach space C[a, b] wasdiscussed earlier in Example 2.9.5, where it was shown that the sequence ofmonomials M = {1, x, x2, . . . } is complete and finitely linearly independentin C[a, b] but is not a Schauder basis for C[a, b].

Page 58: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

54 3 Inner Products and Hilbert Spaces

3.9 Existence of an Orthonormal Basis

A Hilbert space is separable if it contains a countable dense subset. All finite-dimensional Hilbert spaces are separable, and Problem 3.9.20 shows thatthe space ℓ2 is separable. However, not every Hilbert space is separable; anexample is given in Problem 3.9.22.

We will show that every separable Hilbert space contains an orthonormalbasis. We begin with finite-dimensional spaces, where we can use the sameGram–Schmidt procedure that is employed to construct orthonormal se-quences in Rd or Cd.

Theorem 3.9.1. If H is a finite-dimensional Hilbert space and d is the vectorspace dimension of H, then H contains an orthonormal basis of the form{e1, . . . , ed}.

Proof. Since H is a d-dimensional vector space, it has a Hamel basis B thatconsists of d vectors, say B = {x1, . . . , xd}.

Set y1 = x1, and note that f1 6= 0 since x1, . . . , xd are linearly independentand therefore nonzero. Define

M1 = span{x1} = span{y1}.

Note that M1 is closed since it is finite-dimensional.If d = 1 then we stop here. Otherwise M1 is a proper subspace of H, and

x2 /∈ M1 (because {x1, . . . , xd} is linearly independent). Let p2 be the orthog-onal projection of x2 onto M1. Then the vector y2 = x2 − p2 is orthogonal tox1, and y2 6= 0 since x2 /∈ M1. Therefore we can define

M2 = span{x1, x2} = span{y1, y2},

where the second equality follows from the fact that y1, y2 are linear combi-nations of x1, x2, and vice versa.

If d = 2 then we stop here. Otherwise we continue in the same manner.At stage k we have constructed orthogonal vectors y1, . . . , yk such that

Mk = span{x1, . . . , xk} = span{y1, . . . , yk}.

The subspace Mk has dimension k. Consequently, when we reach k = d wewill have Md = H, and therefore

H = Md = span{x1, . . . , xd} = span{y1, . . . , yd}.

The vectors y1, . . . , yd are orthogonal and nonzero, so by setting

ek =yk

‖yk‖, k = 1, . . . , d,

we obtain an orthonormal basis {e1, . . . , ed} for H. ⊓⊔

Page 59: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.9 Existence of an Orthonormal Basis 55

Next we consider infinite-dimensional, but still separable, Hilbert spaces.

Theorem 3.9.2. If H is a infinite-dimensional, separable Hilbert space, thenH contains an orthonormal basis of the form {en}n∈N.

Proof. Since H is separable, it contains a countable dense subset. This subsetmust be infinite, so let us say that it is Z = {zn}n∈N. However, Z need not belinearly independent, so we will extract a linearly independent subsequenceas follows.

Let k1 be the first index such that zk16= 0, and set x1 = zk1

. Then let k2

be the first index larger than k1 such that zk2/∈ span{x1}, and set x2 = zk2

.Then let k3 be the first index larger than k2 such that zk3

/∈ span{x1, x2},and so forth. In this way we obtain vectors x1, x2, . . . such that x1 6= 0 andfor each n > 1 we have

xn /∈ span{x1, . . . , xn−1} and span{x1, . . . , xn} = span{z1, . . . , zkn}.

Therefore {xn}n∈N is linearly independent, and furthermore span(

{xn}n∈N

)

is dense in H.Now we apply the Gram–Schmidt procedure utilized in the proof of Theo-

rem 3.9.1, but without stopping. This gives us orthonormal vectors e1, e2, . . .such that for every n we have

span{e1, . . . , en} = span{x1, . . . , xn}.

Consequently span(

{en}n∈N

)

is dense in H. Therefore {en}n∈N is a completeorthonormal sequence in H, and hence it is an orthonormal basis for H. ⊓⊔

The following result gives a converse to Theorems 3.9.1 and 3.9.2: only aseparable Hilbert space can contain an orthonormal basis. This is a conse-quence of the fact that we declared in Definition 3.8.2 that an orthonormalbasis must be a countable sequence.

Theorem 3.9.3. If a Hilbert space H is not separable, then H does not con-tain an orthonormal basis.

Proof. We will prove the contrapositive statement. Suppose that H containsan orthonormal basis. Such a basis is either finite or countably infinite; sinceboth cases are similar let us assume that {en}n∈N is an orthonormal basisfor H.

Say that a complex number is rational if both its real and imaginary partsare rational, and let

S =

{ N∑

n=1

rnen : N > 0, r1, . . . , rN rational

}

.

This is a countable subset of H, and we will show that it is dense.

Page 60: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

56 3 Inner Products and Hilbert Spaces

Choose any x ∈ H and fix ε > 0. Since ‖x‖2 =∑

|〈x, en〉|2, we can choose

N large enough that∞∑

n=N+1

|〈x, en〉|2 <

ε2

2.

For each n = 1, . . . , N, choose a scalar rn that has real and imaginary partsand satisfies

|〈x, en〉 − rn|2 <

ε2

2N.

Then the vector

z =

N∑

n=1

rnen

belongs to S, and by the Plancherel Equality we have

‖x − z‖2 =

N∑

n=1

|〈x, en〉 − rn|2 +

∞∑

n=N+1

|〈x, en〉|2 < N

ε2

2N+

ε2

2= ε2.

Thus ‖x − z‖ < ε, so S is dense in H. As S is also countable, it follows thatH is separable. ⊓⊔

Nonseparable Hilbert spaces do exist (see Problem 3.9.22 for one example).An argument based on the Axiom of Choice in the form of Zorn’s Lemmashows that every Hilbert space, including nonseparable Hilbert spaces, con-tains a complete orthonormal set (for a proof, see [Heil11, Thm. 1.56]). How-ever, if H is nonseparable, then such a complete orthonormal set must beuncountable. An uncountable complete orthonormal set does have certainbasis-like properties (e.g., see [Heil11, Exer. 3.6]), and for this reason someauthors refer to a complete orthonormal set of any cardinality as an ortho-normal basis. In keeping with the majority of the Banach space literature,we prefer to reserve the word “basis” for use in conjunction with countablesequences only.

Problems

3.9.4. Prove Lemma 3.2.1.

3.9.5. Prove the continuity of the inner product : If H is an inner productspace and xn → x, yn → y in H, then 〈xn, yn〉 → 〈x, y〉.

3.9.6. Prove that if a series∑∞

n=1 xn converges in an inner product space H,then

⟨ ∞∑

n=1

xn, y

=∞∑

n=1

〈xn, y〉, y ∈ H.

Page 61: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.9 Existence of an Orthonormal Basis 57

Note that this is not merely a consequence of the linearity of the inner productin the first variable; the continuity of the inner product is also needed.

3.9.7. Extend the Pythagorean Theorem to finite orthogonal sets of vectors.That is, prove that if x1, . . . , xN ∈ H are orthogonal vectors in an innerproduct space H, then

N∑

n=1

xn

2

=

N∑

n=1

‖xn‖2.

Does the result still hold if we only assume that 〈·, ·〉 is a semi-inner product?

3.9.8. Prove that the function 〈·, ·〉 defined in equation (3.6) is an inner prod-uct on C[a, b].

3.9.9. Prove that ℓ2 is complete with respect to the norm ‖ · ‖2 defined inequation (3.2).

3.9.10. For each n ∈ N, let

yn =(

1, 12 , . . . , 1

n , 0, 0, . . .)

.

Note that yn ∈ ℓ1 for every n.

(a) Assume that the norm on ℓ1 is its usual norm ‖·‖1. Prove that {yn}n∈N

is not a Cauchy sequence in ℓ1 with respect to this norm.

(b) Now assume that the norm on ℓ1 is ‖ · ‖2. Prove that {yn}n∈N is aCauchy sequence in ℓ1 with respect to ‖ · ‖2. Even so, prove that there is novector y ∈ ℓ1 such that ‖y−yn‖2 → 0. Conclude that ℓ1 is not complete withrespect to the norm ‖ · ‖2.

3.9.11. Suppose that {xn}n∈N is a sequence in a Hilbert space H, and y ∈ His orthogonal to xn for every n. Prove the following statements.

(a) y is orthogonal to every vector in span{xn}n∈N, and therefore y ∈

span{xn}n∈N

⊥.

(b) y is orthogonal to every vector in span{xn}n∈N, and therefore y ∈

span{xn}n∈N

⊥.

3.9.12. Prove Lemma 3.5.2.

3.9.13. Prove the remaining implications in Lemma 3.6.3.

3.9.14. Given a sequence {xn}n∈N in a Hilbert space H, prove that the fol-lowing two statements are equivalent.

(a) For each m ∈ N we have xm /∈ span(

{xn}n6=m

)

(such a sequence issaid to be minimal).

Page 62: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

58 3 Inner Products and Hilbert Spaces

(b) There exists a sequence {yn}n∈N in H such that 〈xm, yn〉 = δmn forall m, n ∈ N (we say that sequences {xn}n∈N and {yn}n∈N satisfying thiscondition are biorthogonal).

Show further that in case statements (a), (b) hold, the sequence {yn}n∈N

is unique if and only if {xn}n∈N is complete.

3.9.15. Formulate and prove analogues of Theorems 3.7.1 and 3.8.1 for finiteorthonormal sequences.

3.9.16. A d × d matrix A with scalar entries is said to be positive definite ifAx · x > 0 for all nonzero vectors x ∈ Cd, where x · y denotes the usual dotproduct of vectors in Cd.

(a) Suppose that S is an invertible d×d matrix and Λ is a diagonal matrixwhose diagonal entries are all positive. Prove that A = SΛSH is a positive

definite matrix, where SH = ST is the complex conjugate of the transpose ofS (usually referred to as the Hermitian of S).

Remark: In fact, it can be shown that every positive definite matrix hasthis form, but that is not needed for this problem.

(b) Show that if A is a positive definite d × d matrix, then

〈x, y〉A = Ax · y, x, y ∈ Cd,

defines an inner product on Cd.

(c) Show that if 〈·, ·〉 is an inner product on Cd, then there exists somepositive definite d × d matrix A such that 〈·, ·〉 = 〈·, ·〉A.

3.9.17. Let 〈·, ·〉 be a semi-inner product on a vector space H. Show thatequality holds in the Cauchy–Bunyakovski–Schwarz Inequality if and only ifthere exist scalars α, β, not both zero, such that ‖αx+βy‖ = 0. In particular,if 〈·, ·〉 is an inner product, then either x = cy or y = cx where c is a scalar.

3.9.18. Let M be a closed subspace of a Hilbert space H, and let P bethe orthogonal projection of H onto M. Show that I − P is the orthogonalprojection of H onto M⊥.

3.9.19. Assume {en}n∈N is an orthonormal basis for a Hilbert space H.

(a) Suppose that vectors yn ∈ H satisfy∑

‖en − yn‖2 < 1. Prove that

{yn}n∈N is a complete sequence in H.

(b) Show that part (a) can fail if we only have∑

‖xn − yn‖2 = 1.

3.9.20. Prove that the set S defined in Problem 2.9.13 is a countable, densesubset of ℓ2. Conclude that ℓ2 is separable.

3.9.21. We say that a sequence {xn}n∈N in a Banach space X is ω-dependent

if there exist scalars cn, not all zero, such that∑∞

n=1 cnxn = 0, where the

Page 63: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

3.9 Existence of an Orthonormal Basis 59

series converges in the norm of X. A sequence is ω-independent if it is notω-dependent.

(a) Prove that every Schauder basis is both complete and ω-independent.

(b) Let α, β ∈ C be fixed nonzero scalars such that |α| > |β|. Let {δn}n∈N

be the sequence of standard basis vectors, and define

x0 = δ1 and xn = αδn + βδn+1, n ∈ N.

Prove that the sequence {xn}n≥0 is complete and finitely linearly independentin ℓ2, but it is not ω-independent and therefore is not a Schauder basis for ℓ2.

3.9.22. Let I be an uncountable index set I, and let ℓ2(I) consist of allsequences x = (xi)i∈I with at most countably many nonzero components suchthat

‖x‖22 =

i∈I

|xi|2 < ∞.

(a) Prove that ℓ2(I) is a Hilbert space with respect to the inner product

〈x, y〉 =∑

i∈I

xi yk.

(b) For each i ∈ I, define δi = (δij)j∈I , where δij is the Kronecker delta.Show that {δi}i∈I is a complete orthonormal sequence in ℓ2(I).

(c) Prove that ℓ2(I) is not separable.

Page 64: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric
Page 65: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Index of Symbols

Sets

Symbol Description

∅ Empty set

Br(x) Open ball of radius r centered at x

C Complex plane

N Natural numbers, {1, 2, 3, . . . }

P Set of all polynomials

Q Rational numbers

R Real line

Z Integers, {. . . ,−1, 0, 1, . . . }

Sets

Symbol Description

AC = X\A Complement of a set A ⊆ X

A◦ Interior of a set A

A Closure of a set A

∂A Boundary of a set A

A × B Cartesian product of A and B

dist(x,E) Distance from a point to a set

inf(S) Infimum of a set of real numbers S

span(A) Finite linear span of a set A

span(A) Closure of the finite linear span of A

sup(S) Supremum of a set of real numbers S

61

Page 66: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

62 Symbols

Sequences

Symbol Description

{xn}n∈N A sequence of points x1, x2, . . .

(xk)k∈N A sequence of scalars x1, x2, . . .

δn nth standard basis vector

Functions

Symbol Description

f : A → B A function from A to B

f(A) Direct image of A under f

f−1(B) Inverse image of B under f

Vector Spaces

Symbol Description

c00 Set of all “finite sequences”

Cb(X) Set of bounded continuous functions on X

C[a, b] Set of continuous functions on [a, b]

ℓ1 Set of all absolutely summable sequences

ℓ2 Set of all square-summable sequences

Metrics, Norms, and Inner Products

Symbol Description

A⊥ Orthogonal complement of a set A

d(·, ·) Generic metric

〈·, ·〉 Generic inner product

‖ · ‖ Generic norm

‖f‖u Uniform norm of a function f

x ⊥ y Orthogonal vectors

‖x‖1 ℓ1-norm of a sequence x

‖x‖2 ℓ2-norm of a sequence x

Page 67: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

References

[Ben97] J. J. Benedetto, Harmonic Analysis and Applications, CRC Press, Boca Raton,

FL, 1997.[Con90] J. B. Conway, A Course in Functional Analysis, Second Edition, Springer-

Verlag, New York, 1990.[DM72] H. Dym and H. P. McKean, Fourier Series and Integrals, Academic Press, New

York–London, 1972.[Enf73] P. Enflo, A counterexample to the approximation problem in Banach spaces,

Acta Math., 130 (1973), pp. 309–317.

[Fol99] G. B. Folland, Real Analysis, Second Edition, Wiley, New York, 1999.[Heil11] C. Heil, A Basis Theory Primer, Expanded Edition, Birkhauser, Boston, 2011.[Heil16] C. Heil, A Short Introduction to Metrics, Banach Spaces, and Hilbert Spaces,

Birkhauser, Boston, to appear.

[Kat04] Y. Katznelson, An Introduction to Harmonic Analysis, Third Edition, Cam-bridge University Press, Cambridge, UK, 2004.

[Kre78] E. Kreyszig, Introductory Functional Analysis with Applications, Wiley, New

York, 1978.[Mun75] J. R. Munkres, Topology: A First Course, Prentice-Hall, Englewood Cliffs, NJ,

1975.[Rud76] W. Rudin, Principles of Mathematical Analysis, Third Edition. McGraw-Hill,

New York, 1976.[ST76] I. M. Singer and J. A. Thorpe, Lecture Notes on Elementary Topology and Geom-

etry, Springer-Verlag, New York-Heidelberg, 1976 (reprint of the 1967 edition).[SS05] E. M. Stein and R. Shakarchi, Real Analysis, Princeton University Press, Prince-

ton, NJ, 2005.[WZ77] R. L. Wheeden and A. Zygmund, Measure and Integral, Marcel Dekker, New

York-Basel, 1977.

63

Page 68: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric
Page 69: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

Index

absolutely summable, 2

accumulation point, 7

Baire Category Theorem, 28

Banach space, 23

basis

Hamel, 27

orthonormal, 53

Schauder, 28

Basis Problem, 29

Bessel’s Inequality, 49

biorthogonal sequence, 58

boundary, 7

point, 7

bounded set, 7

Cauchy sequence, 3

Cauchy–Bunyakovski–Schwarz Inequality,38

CBS Inequality, 38

closed

finite span, 26

set, 7

Closest Point Theorem, 45

closure, 7

compact set, 9

complete

inner product space, 39

metric space, 6

normed space, 23

sequence, 27

continuity, 13

of the inner product, 56

of the norm, 22

uniform, 14

convergence

absolute, 24

uniform, 29

convergent

sequence, 3

series, 24

convex set, 23

δ sequence, 3

dense set, 7

diameter, 17

discrete metric, 16

distance, 2, 21

Dvoretzky–Rogers Theorem, 26

finite

linear independence, 28

linear span, 26

function

continuous, 13

uniformly continuous, 14

fundamental sequence, 27

Gram–Schmidt orthogonalization, 54

Hamel basis, 27

Hausdorff metric space, 10, 16

Hermitian, 58

Hilbert space, 39

independence

finite linear, 28

ω-independence, 59

induced

metric, 22

norm, 38

seminorm, 38

inner

product, 37

65

Page 70: A Brief Guide to Metrics, Norms, and Inner Productspeople.math.gatech.edu/~heil/books/metricbrief.pdfA metric space X does not have to be a vector space, although most of the metric

66 Index

product space, 37

interior, 7

Kronecker delta, 44

Lebesgue integral, 43length, 21line segment, 23linear

independence, 28space, 1

metric, 1discrete, 16induced, 22

metric space, 2complete, 6

minimal sequence, 57

norm, 21induced, 38

uniform, 29normalized vector, 21

normed space, 21

complete, 23

openball, 7

set, 7orthogonal

complement, 44

projection, 46

vectors, 44

orthonormalbasis, 53

vectors, 44

Parallelogram Law, 38

Parseval Equality, 52

partial sums, 24

Plancherel Equality, 52

Polar Identity, 38

polynomial, 32

positive definite matrix, 58

power series, 33

pre-Hilbert space, 37

projection

orthogonal, 46Pythagorean Theorem, 38, 57

Reverse Triangle Inequality, 16, 22

scalar, 1Schauder basis, 28Schwarz Inequality, 38

semi-inner product, 37

seminorm, 21

induced, 38

separable space, 7

sequence

absolutely summable, 2

biorthogonal, 58

Cauchy, 3

complete, 27

convergent, 3

fundamental, 27

minimal, 57

square summable, 40

total, 27

ω-independent, 58

sequentially compact, 10

series

absolutely convergent, 24

convergent, 24

unconditionally convergent, 26

set

compact, 9

dense, 7

sequentially compact, 10

totally bounded, 10

span

closed, 26

finite, 26

square summable, 40

standard

basis, 3

metric, 4

summable sequence, 2

topology, 7

total sequence, 27

totally bounded set, 10

Triangle Inequality, 2, 21

unconditionally convergent series, 26

uniform

continuity, 14

convergence, 29

norm, 29

uniformly Cauchy, 29

unit vector, 21

Urysohn’s Lemma, 15

vector

normalized, 21

space, 1

unit, 21

Weierstrass Approximation Theorem, 32