Introduction To Functional Analysis - Daniel Daners - 2017
Introduction To Functional Analysis - Daniel Daners - 2017
Daniel Daners
I Preliminary Material 1
1 The Axiom of Choice and Zorn’s Lemma . . . . . . . . . . . . . . . 1
2 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3 Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4 Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
5 Continuous Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 12
II Banach Spaces 17
6 Normed Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
7 Examples of Banach Spaces . . . . . . . . . . . . . . . . . . . . . . 20
7.1 Elementary Inequalities . . . . . . . . . . . . . . . . . . . . 20
7.2 Spaces of Sequences . . . . . . . . . . . . . . . . . . . . . . 23
7.3 Lebesgue Spaces . . . . . . . . . . . . . . . . . . . . . . . . 25
7.4 Spaces of Bounded and Continuous Functions . . . . . . . . . 26
8 Basic Properties of Bounded Linear Operators . . . . . . . . . . . . . 27
9 Equivalent Norms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
10 Finite Dimensional Normed Spaces . . . . . . . . . . . . . . . . . . 33
11 Infinite Dimensional Normed Spaces . . . . . . . . . . . . . . . . . . 34
12 Quotient Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
IV Hilbert Spaces 47
15 Inner Product Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 47
16 Projections and Orthogonal Complements . . . . . . . . . . . . . . . 52
17 Orthogonal Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
18 Abstract Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . 62
V Linear Operators 69
19 Baire’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
20 The Open Mapping Theorem . . . . . . . . . . . . . . . . . . . . . . 70
21 The Closed Graph Theorem . . . . . . . . . . . . . . . . . . . . . . . 73
i
22 The Uniform Boundedness Principle . . . . . . . . . . . . . . . . . . 74
23 Closed Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
24 Closable Operators and Examples . . . . . . . . . . . . . . . . . . . 81
VI Duality 85
25 Dual Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
26 The Hahn-Banach Theorem . . . . . . . . . . . . . . . . . . . . . . . 89
27 Reflexive Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
28 Weak convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
29 Dual Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
30 Duality in Hilbert Spaces . . . . . . . . . . . . . . . . . . . . . . . . 97
31 The Lax-Milgram Theorem . . . . . . . . . . . . . . . . . . . . . . . 99
Bibliography 117
ii
Acknowledgement
Thanks to Fan Wu from the 2008 Honours Year for providing an extensive list of mis-
prints.
iii
iv
Chapter I
Preliminary Material
In functional analysis many different fields of mathematics come together. The objects
we look at are vector spaces and linear operators. Hence you need to some basic linear
algebra in general vector spaces. I assume your knowledge of that is sufficient. Second
we will need some basic set theory. In particular, many theorems depend on the axiom
of choice. We briefly discuss that most controversial axiom of set theory and some
equivalent statements. In addition to the algebraic structure on a vector space, we will
look at topologies on them. Of course, these topologies should be compatible with the
algebraic structure. This means that addition and multiplication by scalars should be
continuous with respect to the topology. We will only look at one class of such spaces,
namely normed spaces which are naturally metric spaces. Hence it is essential you
know the basics of metric spaces, and we provide a self contained introduction of what
we need in the course.
such that 𝑥(𝛼) ∈ 𝑋𝛼 for all 𝛼 ∈ 𝐴. We write 𝑥𝛼 for 𝑥(𝛼) and (𝑥𝛼 )𝛼∈𝐴 or simply (𝑥𝛼 )
for a given such function 𝑥. Suppose now that 𝐴 ≠ ∅ and 𝑋𝛼 ≠ ∅ for all 𝛼 ∈ 𝐴. Then
there is a fundamental question:
∏
Is 𝛼∈𝐴 𝑋𝛼 nonempty in general?
Here some brief history about the problem, showing how basic and difficult it is:
1
• Zermelo (1904) (see [14]) observed that it is not obvious from the existing axioms
of set theory that there is a procedure to select a single 𝑥𝛼 from each 𝑋𝛼 in general.
As a consequence he introduced what we call the axiom of choice, asserting that
∏
𝛼∈𝐴 𝑋𝛼 ≠ ∅ whenever 𝐴 ≠ ∅ and 𝑋𝛼 ≠ ∅ for all 𝛼 ∈ 𝐴.
It remained open whether his axiom of choice could be derived from the other
axioms of set theory. There was an even more fundamental question on whether
the axiom is consistent with the other axioms!
• Gödel (1938) (see [8]) proved that the axiom of choice is consistent with the other
axioms of set theory. The open question remaining was whether it is independent
of the other axioms.
• P.J. Cohen (1963/1964) (see [4, 5]) finally showed that the axiom of choice is in
fact independent of the other axioms of set theory, that is, it cannot be derived
from them.
The majority of mathematicians accept the axiom of choice, but there is a minority
which does not. Many very basic and important theorems in functional analysis cannot
be proved without the axiom of choice.
We accept the axiom of choice.
There are some non-trivial equivalent formulations of the axiom of choice which are
useful for our purposes. Given two sets 𝑋 and 𝑌 recall that a relation from 𝑋 to 𝑌 is
simply a subset of the Cartesian product 𝑋 ×𝑌 . We now explore some special relations,
namely order relations.
2
(a) 𝑚 ∈ 𝑋 is called a maximal element in 𝑋 if for all 𝑥 ∈ 𝑋 with 𝑥 ≻ 𝑚 we have
𝑥 ≺ 𝑚;
(d) If a partially ordered set (𝑋, ≺) is a chain we call it a totally ordered set.
(e) If (𝑋, ≺) is partially ordered and 𝑥0 ∈ 𝑋 is such that 𝑥0 ≺ 𝑥 for all 𝑥 ∈ 𝑋, then
we call 𝑥0 a first element.
There is a special class of partially ordered sets playing a particularly important role in
relation to the axiom of choice as we will see later.
1.4 Definition (well ordered set) A partially ordered set (𝑋, ≺) is called a well or-
dered set if every subset has a first element.
1.5 Examples (a) ℕ is a well ordered set, but ℤ or ℝ are not well ordered with the
usual order.
(b) ℤ and ℝ are totally ordered with the usual order.
1.6 Remark Well ordered sets are always totally ordered. To see this assume (𝑋, ≺)
is well ordered. Given 𝑥, 𝑦 ∈ 𝑋 we consider the subset {𝑥, 𝑦} of 𝑋. By definition of
a well ordered set we have either 𝑥 ≺ 𝑦 or 𝑦 ≺ 𝑥, which shows that (𝑋, ≺) is totally
ordered. The converse is not true as the example of ℤ given above shows.
There is another, highly non-obvious but very useful statement appearing in connection
with partially ordered sets:
1.7 Zorn’s Lemma Suppose that (𝑋, ≺) is a partially ordered set such that each chain
in 𝑋 has an upper bound. Then 𝑋 has a maximal element.
There is a non-trivial connection between all the apparently different topics we dis-
cussed so far. We state it without proof (see for instance [7]).
The axiom of choice may seem “obvious” at the first instance. However, the other two
equivalent statements are certainly not. For instance take 𝑋 = ℝ, which we know is not
well ordered with the usual order. If we accept the axiom of choice then it follows from
the above theorem that there exists a partial ordering making ℝ into a well ordered
set. This is a typical “existence proof” based on the axiom of choice. It does not
give us any hint on how to find a partial ordering making ℝ into a well ordered set.
This reflects Zermelo’s observation that it is not obvious how to choose precisely one
3
element from each set when given an arbitrary collection of sets. Because of the non-
constructive nature of the axiom of choice and its equivalent counterparts, there are
some mathematicians rejecting the axiom. These mathematicians have the point of
view that everything should be “constructible,” at least in principle, by some means
(see for instance [2]).
2 Metric Spaces
Metric spaces are sets in which we can measure distances between points. We expect
such a “distance function,” called a metric, to have some obvious properties, which we
postulate in the following definition.
We call (𝑋, 𝑑) a metric space. If it is clear what metric is being used we simply say 𝑋
is a metric space.
2.2 Example The simplest example of a metric space is ℝ with 𝑑(𝑥, 𝑦) ∶= |𝑥 − 𝑦|.
The standard metric used in ℝ𝑁 is the Euclidean metric given by
√
√𝑁
√∑
𝑑(𝒙, 𝒚) = |𝒙 − 𝒚|2 ∶= √ |𝑥𝑖 − 𝑦𝑖 |2
𝑖=1
for all 𝒙, 𝒚 ∈ ℝ𝑁 .
2.3 Remark If (𝑋, 𝑑) is a metric space, then every subset 𝑌 ⊆ 𝑋 is a metric space
with the metric restricted to 𝑌 . We say the metric on 𝑌 is induced by the metric on 𝑋.
2.4 Definition (Open and Closed Ball) Let (𝑋, 𝑑) be a metric space. For 𝑟 > 0 we
call
𝐵(𝑥, 𝑟) ∶= {𝑦 ∈ 𝑋 ∶ 𝑑(𝑥, 𝑦) < 𝑟}
the open ball about 𝑥 with radius 𝑟. Likewise we call
̄ 𝑟) ∶= {𝑦 ∈ 𝑋 ∶ 𝑑(𝑥, 𝑦) ≤ 𝑟}
𝐵(𝑥,
4
2.5 Definition (Open and Closed Set) Let (𝑋, 𝑑) be a metric space. A subset 𝑈 ⊆ 𝑋
is called open if for every 𝑥 ∈ 𝑋 there exists 𝑟 > 0 such that 𝐵(𝑥, 𝑟) ⊆ 𝑈 . A set 𝑈 is
called closed if its complement 𝑋 ⧵ 𝑈 is open.
2.6 Remark For every 𝑥 ∈ 𝑋 and 𝑟 > 0 the open ball 𝐵(𝑥, 𝑟) in a metric space is
open. To prove this fix 𝑦 ∈ 𝐵(𝑥, 𝑟). We have to show that there exists 𝜀 > 0 such that
𝐵(𝑦, 𝜀) ⊆ 𝐵(𝑥, 𝑟). To do so note that by definition 𝑑(𝑥, 𝑦) < 𝑟. Hence we can choose
𝜀 ∈ ℝ such that 0 < 𝜀 < 𝑟 − 𝑑(𝑥, 𝑦). Thus, by property (iv) of a metric, for 𝑧 ∈ 𝐵(𝑦, 𝜀)
we have 𝑑(𝑥, 𝑧) ≤ 𝑑(𝑥, 𝑦) + 𝑑(𝑦, 𝑧) < 𝑑(𝑥, 𝑦) + 𝑟 − 𝑑(𝑥, 𝑦) = 𝑟. Therefore 𝑧 ∈ 𝐵(𝑥, 𝑟),
showing that 𝐵(𝑦, 𝜀) ⊆ 𝐵(𝑥, 𝑟).
Next we collect some fundamental properties of open sets.
2.7 Theorem Open sets in a metric space (𝑋, 𝑑) have the following properties.
(i) 𝑋, ∅ are open sets;
2.8 Remark There is a more general concept than that of a metric space, namely that
of a “topological space.” A collection of subsets of a set 𝑋 is called a topology if the
following conditions are satisfied
(i) 𝑋, ∅ ∈ ;
2.9 Definition (Neighbourhood) Suppose that (𝑋, 𝑑) is a metric space (or more gen-
erally a topological space). We call a set 𝑈 a neighbourhood of 𝑥 ∈ 𝑋 if there exists
an open set 𝑉 ⊆ 𝑈 with 𝑥 ∈ 𝑉 .
Now we define some sets associated with a given subset of a metric space.
5
(i) 𝑈̊ ∶= Int(𝑈 ) ∶= {𝑥 ∈ 𝑈 ∶ 𝑥 interior point of 𝑈 } the interior of 𝑈 ;
2.11 Remark A set is open if and only if 𝑈̊ = 𝑈 and closed if and only if 𝑈 = 𝑈 .
Moreover, 𝜕𝑈 = 𝑈 ∩ 𝑋 ⧵ 𝑈 .
Sometimes it is convenient to look at products of a (finite) number of metric spaces.
It is possible to define a metric on such a product as well.
∑
𝑛
𝑑(𝑥, 𝑦) ∶= 𝑑𝑖 (𝑥𝑖 , 𝑦𝑖 )
𝑖=1
∑
𝑛
∑
𝑛
( )
𝑑(𝑥, 𝑦) = 𝑑(𝑥𝑖 , 𝑦𝑖 ) ≤ 𝑑(𝑥𝑖 , 𝑧𝑖 ) + 𝑑(𝑧𝑖 , 𝑦𝑖 )
𝑖=1 𝑖=1
∑
𝑛
∑
𝑛
= 𝑑(𝑥𝑖 , 𝑧𝑖 ) + 𝑑(𝑧𝑖 , 𝑦𝑖 ) = 𝑑(𝑥, 𝑦) + 𝑑(𝑧, 𝑦)
𝑖=1 𝑖=1
for all 𝑥, 𝑦, 𝑧 ∈ 𝑋.
2.13 Definition (Product space) The space and metric introduced in Proposition 2.12
is called a product space and a product metric, respectively.
3 Limits
Once we have a notion of “closeness” we can discuss the asymptotics of sequences and
continuity of functions.
3.1 Definition (Limit) Suppose (𝑥𝑛 )𝑛∈ℕ is a sequence in a metric space (𝑋, 𝑑), or more
generally a topological space. We say 𝑥0 is a limit of (𝑥𝑛 ) if for every neighbourhood
𝑈 of 𝑥0 there exists 𝑛0 ∈ ℕ such that 𝑥𝑛 ∈ 𝑈 for all 𝑛 ≥ 𝑛0 . We write
𝑥0 = lim 𝑥𝑛 or 𝑥𝑛 → 𝑥 as 𝑛 → ∞.
𝑛→∞
6
3.2 Remark Let (𝑥𝑛 ) be a sequence in a metric space (𝑋, 𝑑) and 𝑥0 ∈ 𝑋. Then the
following statements are equivalent:
(1) lim 𝑥𝑛 = 𝑥0 ;
𝑛→∞
(2) for every 𝜀 > 0 there exists 𝑛0 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑥0 ) < 𝜀 for all 𝑛 ≥ 𝑛0 .
Proof. Clearly (1) implies (2) by choosing neighbourhoods of the form 𝐵(𝑥, 𝜀). If
(2) holds and 𝑈 is an arbitrary neighbourhood of 𝑥0 we can choose 𝜀 > 0 such that
𝐵(𝑥0 , 𝜀) ⊆ 𝑈 . By assumption there exists 𝑛0 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑥0 ) < 𝜀 for all 𝑛 ≥ 𝑛0 ,
that is, 𝑥𝑛 ∈ 𝐵(𝑥0 , 𝜀) ⊆ 𝑈 for all 𝑛 ≥ 𝑛0 . Therefore, 𝑥𝑛 → 𝑥0 as 𝑛 → ∞.
3.3 Proposition A sequence in a metric space (𝑋, 𝑑) has at most one limit.
Proof. Suppose that (𝑥𝑛 ) is a sequence in (𝑋, 𝑑) and that 𝑥 and 𝑦 are limits of that
sequence. Fix 𝜀 > 0 arbitrary. Since 𝑥 is a limit there exists 𝑛1 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑥) <
𝜀∕2 for all 𝑛 > 𝑛1 . Similarly, since 𝑦 is a limit there exists 𝑛2 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑦) <
𝜀∕2 for all 𝑛 > 𝑛2 . Hence 𝑑(𝑥, 𝑦) ≤ 𝑑(𝑥, 𝑥𝑛 ) + 𝑑(𝑥𝑛 , 𝑦) ≤ 𝜀∕2 + 𝜀∕2 = 𝜀 for all
𝑛 > max{𝑛1 , 𝑛2 }. Since 𝜀 > 0 was arbitrary it follows that 𝑑(𝑥, 𝑦) = 0, and so by
definition of a metric 𝑥 = 𝑦. Thus (𝑥𝑛 ) has at most one limit.
3.4 Theorem Let 𝑈 be a subset of the metric space (𝑋, 𝑑) then 𝑥 ∈ 𝑈 if and only if
there exists a sequence (𝑥𝑛 ) in 𝑈 such that 𝑥𝑛 → 𝑥 as 𝑛 → ∞.
Proof. Let 𝑈 ⊆ 𝑋 and 𝑥 ∈ 𝑈 . Hence 𝐵(𝑥, 𝜀) ∩ 𝑈 ) ≠ ∅ for all 𝜀 > 0. For all 𝑛 ∈ ℕ we
can therefore choose 𝑥𝑛 ∈ 𝑈 with 𝑑(𝑥, 𝑥𝑛 ) < 1∕𝑛. By construction 𝑥𝑛 → 𝑥 as 𝑛 → ∞.
If (𝑥𝑛 ) is a sequence in 𝑈 converging to 𝑥 then for every 𝜀 > 0 there exists 𝑛0 ∈ ℕ such
that 𝑥𝑛 ∈ 𝐵(𝑥, 𝜀) for all 𝑛 ≥ 𝑛0 . In particular, 𝐵(𝑥, 𝜀) ∩ 𝑈 ≠ ∅ for all 𝜀 > 0, implying
that 𝑥 ∈ 𝑈 as required.
3.5 Definition (Cauchy Sequence) Suppose (𝑥𝑛 ) is a sequence in the metric space
(𝑋, 𝑑). We call (𝑥𝑛 ) a Cauchy sequence if for every 𝜀 > 0 there exists 𝑛0 ∈ ℕ such that
𝑑(𝑥𝑛 , 𝑥𝑚 ) < 𝜀 for all 𝑚, 𝑛 ≥ 𝑛0 .
Some sequences may not converge, but they accumulate at certain points.
3.7 Remark Equivalently we may say 𝑥0 is an accumulation point of (𝑥𝑛 ) if for every
𝜀 > 0 and every 𝑛0 ∈ ℕ there exists 𝑛 ≥ 𝑛0 such that 𝑑(𝑥𝑛 , 𝑥0 ) < 𝜀. Note that it follows
from the definition that every neighbourhood of 𝑥0 contains infinitely many elements
of the sequence (𝑥𝑛 ).
7
3.8 Proposition Suppose that (𝑋, 𝑑) is a metric space and (𝑥𝑛 ) a sequence in that
space. Then 𝑥 ∈ 𝑋 is a point of accumulation of (𝑥𝑛 ) if and only if
⋂
∞
𝑥∈ {𝑥𝑗 ∶ 𝑗 ≥ 𝑘}. (3.1)
𝑘=1
⋂∞
Proof. Suppose that 𝑥 ∈ 𝑘=1 {𝑥𝑗 ∶ 𝑗 ≥ 𝑘}. Then 𝑥 ∈ {𝑥𝑗 ∶ 𝑗 ≥ 𝑘} for all 𝑘 ∈ ℕ.
By Theorem 3.4 we can choose for every 𝑘 ∈ ℕ an element 𝑥𝑛𝑘 ∈ {𝑥𝑗 ∶ 𝑗 ≥ 𝑘} such
that 𝑑(𝑥𝑛𝑘 , 𝑥) < 1∕𝑘. By construction 𝑥𝑛𝑘 → 𝑥 as 𝑘 → ∞, showing that 𝑥 is a point
of accumulation of (𝑥𝑛 ). If 𝑥 is a point of accumulation of (𝑥𝑛 ) then for all 𝑘 ∈ ℕ
there exists 𝑛𝑘 ≥ 𝑘 such that 𝑑(𝑥𝑛𝑘 , 𝑥) < 1∕𝑘. Clearly 𝑥𝑛𝑘 → 𝑥 as 𝑘 → ∞, so that
𝑥 ∈ {𝑥𝑛𝑗 ∶ 𝑗 ≥ 𝑘} for all 𝑘 ∈ ℕ. As {𝑥𝑛𝑗 ∶ 𝑗 ≥ 𝑘} ⊆ {𝑥𝑗 ∶ 𝑗 ≥ 𝑘} for all 𝑘 ∈ ℕ we
obtain (3.1).
In the following theorem we establish a connection between Cauchy sequences and
converging sequences.
3.9 Theorem Let (𝑋, 𝑑) be a metric space. Then every convergent sequence is a
Cauchy sequence. Moreover, if a Cauchy sequence (𝑥𝑛 ) has an accumulation point
𝑥0 , then (𝑥𝑛 ) is a convergent sequence with limit 𝑥0 .
Proof. Suppose that (𝑥𝑛 ) is a convergent sequence with limit 𝑥0 . Then for every 𝜀 > 0
there exists 𝑛0 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑥0 ) < 𝜀∕2 for all 𝑛 ≥ 𝑛0 . Now
𝜀 𝜀
𝑑(𝑥𝑛 , 𝑥𝑚 ) ≤ 𝑑(𝑥𝑛 , 𝑥0 ) + 𝑑(𝑥0 , 𝑥𝑚 ) = 𝑑(𝑥𝑛 , 𝑥0 ) + 𝑑(𝑥𝑚 , 𝑥0 ) < + = 𝜀
2 2
for all 𝑛, 𝑚 ≥ 𝑛0 , showing that (𝑥𝑛 ) is a Cauchy sequence. Now assume that (𝑥𝑛 ) is a
Cauchy sequence, and that 𝑥0 ∈ 𝑋 is an accumulation point of (𝑥𝑛 ). Fix 𝜀 > 0 arbitrary.
Then by definition of a Cauchy sequence there exists 𝑛0 ∈ ℕ such that 𝑑(𝑥𝑛 , 𝑥𝑚 ) < 𝜀∕2
for all 𝑛, 𝑚 ≥ 𝑛0 . Moreover, since 𝑥0 is an accumulation point there exists 𝑚0 ≥ 𝑛0 such
that 𝑑(𝑥𝑚0 , 𝑥0 ) < 𝜀∕2. Hence
𝜀 𝜀
𝑑(𝑥𝑛 , 𝑥0 ) ≤ 𝑑(𝑥𝑛 , 𝑥𝑚0 ) + 𝑑(𝑥𝑚0 , 𝑥0 ) < + = 𝜀
2 2
for all 𝑛 ≥ 𝑛0 . Hence by Remark 3.2 𝑥0 is the limit of (𝑥𝑛 ).
In a general metric space not all Cauchy sequences have necessarily a limit, hence the
following definition.
3.10 Definition (Complete Metric Space) A metric space is called complete if every
Cauchy sequence in that space has a limit.
One property of the real numbers is that the intersection of a nested sequence of closed
bounded intervals whose lengths shrinks to zero have a non-empty intersection. This
property is in fact equivalent to the “completeness” of the real number system. We now
prove a counterpart of that fact for metric spaces. There are no intervals in general
metric spaces, so we look at a sequence of nested closed sets whose diameter goes to
zero. The diameter of a set 𝐾 in a metric space (𝑋, 𝑑) is defined by
diam(𝐾) ∶= sup 𝑑(𝑥, 𝑦).
𝑥,𝑦∈𝐾
8
3.11 Theorem (Cantor’s Intersection Theorem) Let (𝑋, 𝑑) be a metric space. Then
the following two assertions are equivalent:
(ii) For every sequence of closed sets 𝐾𝑛 ⊆ 𝑋 with 𝐾𝑛+1 ⊆ 𝐾𝑛 for all 𝑛 ∈ ℕ
⋂
as 𝑛 → ∞ we have 𝑛∈ℕ 𝐾𝑛 ≠ ∅.
Proof. First assume that 𝑋 is complete and let 𝐾𝑛 be as in (ii). For every 𝑛 ∈ ℕ we
choose 𝑥𝑛 ∈ 𝐾𝑛 and show that (𝑥𝑛 ) is a Cauchy sequence. By assumption 𝐾𝑛+1 ⊆ 𝐾𝑛
for all 𝑛 ∈ ℕ, implying that 𝑥𝑚 ∈ 𝐾𝑚 ⊆ 𝐾𝑛 for all 𝑚 > 𝑛. Since 𝑥𝑚 , 𝑥𝑛 ∈ 𝐾𝑛 we have
for all 𝑚 > 𝑛. Since diam(𝐾𝑛 ) → 0 as 𝑛 → ∞, given 𝜀 > 0 there exists 𝑛0 ∈ ℕ such
that diam(𝐾𝑛0 ) < 𝜀. Hence, since 𝐾𝑚 ⊆ 𝐾𝑛 ⊆ 𝐾𝑛0 we have
for all 𝑚 > 𝑛 > 𝑛0 , showing that (𝑥𝑛 ) is a Cauchy sequence. By completenes of 𝑆,
the sequence (𝑥𝑛 ) converges to some 𝑥 ∈ 𝑋. We know from above that 𝑥𝑚 ∈ 𝐾𝑛 for
all 𝑚 > 𝑛. As 𝐾𝑛 is closed 𝑥 ∈ 𝐾𝑛 . Since this is true for all 𝑛 ∈ ℕ we conclude that
⋂
𝑥 ∈ 𝑛∈ℕ 𝐾𝑛 , so the intersection is non-empty as claimed.
Assume now that (ii) is true and let (𝑥𝑛 ) be a Cauchy sequence in (𝑋, 𝑑). Hence
there exists 𝑛0 ∈ 𝑋 such that 𝑑(𝑥𝑛0 , 𝑥𝑛 ) < 1∕2 for all 𝑛 ≥ 𝑛0 . Similarly, there exists
𝑛1 > 𝑛0 such that 𝑑(𝑥𝑛1 , 𝑥𝑛 ) < 1∕22 for all 𝑛 ≥ 𝑛1 . Continuing that way we construct a
sequence (𝑛𝑘 ) in ℕ such that for every 𝑘 ∈ ℕ we have 𝑛𝑘+1 > 𝑛𝑘 and 𝑑(𝑥𝑛𝑘 , 𝑥𝑛 ) < 1∕2𝑘+1
for all 𝑛 > 𝑛𝑘 . We now set 𝐾𝑘 ∶= 𝐵(𝑥𝑘 , 2−𝑘 )). If 𝑥 ∈ 𝐾𝑘+1 , then since 𝑛𝑘+1 > 𝑛𝑘
1 1 1
𝑑(𝑥𝑛𝑘 , 𝑥) ≤ 𝑑(𝑥𝑛𝑘 , 𝑥𝑛𝑘+1 ) + 𝑑(𝑥𝑛𝑘+1 , 𝑥) < + = .
2𝑘+1 2𝑘+1 2𝑘
Hence 𝑥 ∈ 𝐾𝑘 , showing that 𝐾𝑘+1 ⊆ 𝐾𝑘 for all 𝑘 ∈ ℕ. By assumption (ii) we have
⋂ ⋂
𝑘∈ℕ 𝐾𝑘 ≠ ∅, so choose 𝑥 ∈ 𝑘∈ℕ 𝐾𝑘 ≠ ∅. Then 𝑥 ∈ 𝐾𝑘 for all 𝑘 ∈ ℕ, so 𝑑(𝑥𝑛𝑘 , 𝑥) ≤
1∕2 for all 𝑘 ∈ ℕ. Hence 𝑥𝑛𝑘 → 𝑥 as 𝑘 → ∞. By Theorem 3.9 the Cauchy sequence
𝑘
We finally look at product spaces defined in Definition 2.13. The rather simple
proof of the following proposition is left to the reader.
3.12 Proposition Suppose that (𝑋𝑖 , 𝑑𝑖 ), 𝑖 = 1, … , 𝑛 are complete metric spaces. Then
the corresponding product space is complete with respect to the product metric.
9
4 Compactness
We start by introducing some additional concepts, and show that they are all equivalent
in a metric space. They are all generalisations of “finiteness” of a set.
4.1 Definition (Open Cover, Compactness) Let (𝑋, 𝑑) be a metric space. We call a
⋃
collection of open sets (𝑈𝛼 )𝛼∈𝐴 an open cover of 𝑋 if 𝑋 ⊆ 𝛼∈𝐴 𝑈𝛼 . The space 𝑋
is called compact if for every open cover (𝑈𝛼 )𝛼∈𝐴 there exist finitely many 𝛼𝑖 ∈ 𝐴,
𝑖 = 1, … , 𝑚 such that (𝑈𝛼𝑖 )𝑖=1,…,𝑚 is an open cover of 𝑋. We talk about a finite sub-
cover of 𝑋.
4.3 Definition (Total Boundedness) We call a metric space 𝑋 totally bounded if for
every 𝜀 > 0 there exist finitely many points 𝑥𝑖 ∈ 𝑋, 𝑖 = 1, … , 𝑚, such that (𝐵(𝑥𝑖 , 𝜀))𝑖=1,…,𝑚
is an open cover of 𝑋.
It turns out that all the above definitions are equivalent, at least in metric spaces (but
not in general topological spaces).
4.4 Theorem For a metric space (𝑋, 𝑑) the following statements are equivalent:
(i) 𝑋 is compact;
which is equivalent to
⋃ ⋃ ⋂
𝑈𝑛 = 𝑋 ⧵ 𝐶𝑛 = 𝑋 ⧵ 𝐶𝑛 ≠ 𝑋
𝑛∈ℕ 𝑛∈ℕ 𝑛∈ℕ
10
we may construct a sequence (𝑥𝑛 ) such that 𝑑(𝑥𝑗 , 𝑥𝑛 ) ≥ 𝜀 for all 𝑗 = 1, … , 𝑛 − 1.
Indeed, suppose we have 𝑥0 , … , 𝑥𝑛 ∈ 𝑋 with 𝑑(𝑥𝑗 , 𝑥𝑛 ) ≥ 𝜀 for all 𝑗 = 1, … , 𝑛 − 1.
⋃𝑛
Assuming that 𝑋 is not totally bounded 𝑗=1 𝐵(𝑥𝑗 , 𝜀) ≠ 𝑋, so we can choose 𝑥𝑛+1
not in that union. Hence 𝑑(𝑥𝑗 , 𝑥𝑛+1 ) ≥ 𝜀 for 𝑗 = 1, … , 𝑛. By construction it follows
that 𝑑(𝑥𝑛 , 𝑥𝑚 ) ≥ 𝜀∕2 for all 𝑛, 𝑚 ∈ ℕ, showing that (𝑥𝑛 ) does not contain a Cauchy
subsequence, and thus has no point of accumulation. As this contradicts (ii), the space
𝑋 must be totally bounded.
Suppose now that (iii) holds, but 𝑋 is not compact. Then there exists an open cover
(𝑈𝛼 )𝛼∈𝐴 not having a finite sub-cover. As 𝑋 is totally bounded, for every 𝑛 ∈ ℕ there
exist finite sets 𝐹𝑛 ⊆ 𝑋 such that
⋃
𝑋= 𝐵(𝑥, 2−𝑛 ). (4.1)
𝑥∈𝐹𝑛
Assuming that (𝑈𝛼 )𝛼∈𝐴 does not have a finite sub-cover, there exists 𝑥1 ∈ 𝐹1 such
that 𝐵(𝑥1 , 2−1 ) and thus 𝐾1 ∶= 𝐵(𝑥1 , 3 ⋅ 2−1 ) cannot be covered by finitely many 𝑈𝛼 .
By (4.1) it follows that there exists 𝑥2 ∈ 𝐹2 such that 𝐵(𝑥1 , 2−1 ) ∩ 𝐵(𝑥2 , 2−2 ) and
therefore 𝐾2 ∶= 𝐵(𝑥2 , 3 ⋅ 2−2 ) is not finitely covered by (𝑈𝛼 )𝛼∈𝐴 . We can continue
this way and choose 𝑥𝑛+1 ∈ 𝐹𝑛+1 such that 𝐵(𝑥𝑛 , 2−𝑛 ) ∩ 𝐵(𝑥𝑛+1 , 2−(𝑛+1) ) and therefore
𝐾𝑛+1 ∶= 𝐵(𝑥2 , 3 ⋅ 2−(𝑛+1) ) is not finitely covered by (𝑈𝛼 )𝛼∈𝐴 . Note that 𝐵(𝑥𝑛 , 2−𝑛 ) ∩
𝐵(𝑥𝑛+1 , 2−(𝑛+1) ) ≠ ∅ since otherwise the intersection is finitely covered by (𝑈𝛼 )𝛼∈𝐴 .
Hence if 𝑥 ∈ 𝐾𝑛+1 , then
1 1 3 6 3
𝑑(𝑥𝑛 , 𝑥) ≤ 𝑑(𝑥𝑛 , 𝑥𝑛+1 ) + 𝑑(𝑥𝑛+1 , 𝑥) ≤ 𝑛
+ 𝑛+1 + 𝑛+1 = 𝑛+1 = 𝑛 ,
2 2 2 2 2
implying that 𝑥 ∈ 𝐾𝑛 . Also diam 𝐾𝑛 ≤ 3 ⋅ 2𝑛−1 → 0. Since 𝑋 is complete, by
⋂
Cantor’s intersection Theorem 3.11 there exists 𝑥 ∈ 𝑛∈ℕ 𝐾𝑛 . As (𝑈𝛼 ) is a cover of
𝑋 we have 𝑥 ∈ 𝑈𝛼0 for some 𝛼0 ∈ 𝐴. Since 𝑈𝛼0 is open there exists 𝜀 > 0 such that
𝐵(𝑥, 𝜀) ⊆ 𝑈𝛼0 . Choose now 𝑛 such that 6∕2𝑛 < 𝜀 and fix 𝑦 ∈ 𝐾𝑛 . Since 𝑥 ∈ 𝐾𝑛 we
have 𝑑(𝑥, 𝑦) ≤ 𝑑(𝑥, 𝑥𝑛 ) + 𝑑(𝑥𝑛 , 𝑦) ≤ 6∕2𝑛 < 𝜀. Hence 𝐾𝑛 ⊆ 𝐵(𝑥, 𝜀) ⊆ 𝑈𝛼0 , showing
that 𝐾𝑛 is covered by 𝑈𝛼0 . However, by construction 𝐾𝑛 cannot be covered by finitely
many 𝑈𝛼 , so we have a contradiction. Hence 𝑋 is compact, completing the proof of the
theorem.
The last part of the proof is modelled on the usual proof of the Heine-Borel theorem
asserting that bounded and closed sets are the compact sets in ℝ𝑁 . Hence it is not a
surprise that the Heine-Borel theorem easily follows from the above characterisations
of compactness.
11
that 𝐴 is contained in the cube [−𝑀, 𝑀]𝑁 . Given 𝜀 > 0 the interval [−𝑀, 𝑀] can
be covered by 𝑚 ∶= [2𝑀∕𝜀] + 1 closed intervals of length 𝜀∕2 (here [2𝑀∕𝜀] is the
integer part of 2𝑀∕𝜀). Hence [−𝑀, 𝑀]𝑁 can be covered by 𝑚𝑁 cubes with edges 𝜀∕2
long. Such cubes are contained in open balls of radius 𝜀, so we can cover [−𝑀, 𝑀]𝑁
and thus 𝐴 by a finite number of balls of radius 𝜀. Hence 𝐴 is complete and totally
bounded. By Theorem 4.4 the set 𝐴 is compact.
We can also look at subsets of metric spaces. As they are metric spaces with the metric
induced on them we can talk about compact subsets of a metric space. It follows from
the above theorem that compact subsets of a metric space are always closed (as they are
complete). Often in applications one has sets that are not compact, but their closure is
compact.
4.6 Definition (Relatively Compact Sets) We call a subset of a metric space relatively
compact if its closure is compact.
Next we show that finite products of compact metric spaces are compact.
4.8 Proposition Let (𝑋𝑖 , 𝑑𝑖 ), 𝑖 = 1, … , 𝑛, be compact metric spaces. Then the prod-
uct 𝑋 ∶= 𝑋1 × ⋯ × 𝑋𝑛 is compact with respect to the product metric introduced in
Proposition 2.12.
Proof. By Proposition 3.12 it follows that the product space 𝑋 is complete. By
Theorem 4.4 is is therefore sufficient to show that 𝑋 is totally bounded. Fix 𝜀 > 0.
Since 𝑋𝑖 is totally bounded there exist 𝑥𝑖𝑘 ∈ 𝑋𝑖 , 𝑘 = 1, … 𝑚𝑖 such that 𝑋𝑖 is covered
by the balls 𝐵𝑖𝑘 of radius 𝜀∕𝑛 and centre 𝑥𝑖𝑘 . Then 𝑋 is covered by the balls of ra-
dius 𝜀 with centres (𝑥1𝑘1 , … , 𝑥𝑖𝑘𝑖 , … 𝑥𝑛𝑘𝑛 ), where 𝑘𝑖 = 1, … 𝑚𝑖 . Indeed, suppose that
𝑥 = (𝑥1 , 𝑥2 , … , 𝑥𝑛 ) ∈ 𝑋 is arbitrary. By assumption, for every 𝑖 = 1, … 𝑛 there exist
1 ≤ 𝑘𝑖 ≤ 𝑚𝑖 such that 𝑑(𝑥𝑖 , 𝑥𝑖𝑘𝑖 ) < 𝜀∕𝑛. By definition of the product metric the distance
between (𝑥1𝑘1 , … , 𝑥𝑛𝑘𝑛 ) and 𝑥 is no larger than 𝑑(𝑥1 , 𝑥1𝑘1 )+⋯+𝑑(𝑥𝑛 , 𝑥𝑛𝑘𝑛 ) ≤ 𝑛𝜀∕𝑛 = 𝜀.
Hence 𝑋 is totally bounded and thus 𝑋 is compact.
5 Continuous Functions
We give a brief overview on continuous functions between metric spaces. Throughout,
let 𝑋 = (𝑋, 𝑑) denote a metric space. We start with some basic definitions.
12
there exists a neighbourhood 𝑈 ⊆ 𝑋 of 𝑥 such that 𝑓 (𝑈 ) ⊆ 𝑉 . The map 𝑓 ∶ 𝑋 → 𝑌
is called continuous if it is continuous at all 𝑥 ∈ 𝑋. Finally we set
𝐶(𝑋, 𝑌 ) ∶= {𝑓 ∶ 𝑋 → 𝑌 ∣ 𝑓 is continuous}.
5.2 Theorem Let 𝑋, 𝑌 be metric spaces and 𝑓 ∶ 𝑋 → 𝑌 a function. Then the follow-
ing assertions are equivalent:
(i) 𝑓 is continuous at 𝑥 ∈ 𝑋;
( )
(ii) For every 𝜀 > 0 there exists 𝛿 > 0 such that 𝑑𝑌 𝑓 (𝑥), 𝑓 (𝑦) ≤ 𝜀 for all 𝑦 ∈ 𝑋
with 𝑑𝑋 (𝑥, 𝑦) < 𝛿;
Proof. Taking special neighbourhoods 𝑉 = 𝐵(𝑓 (𝑥), 𝜀) and 𝑈 ∶= 𝐵(𝑥, 𝛿) then (ii)
is clearly necessary for 𝑓 to be continuous. To show the (ii) is sufficient let 𝑉 be an
arbitrary neighbourhood of 𝑓 (𝑥). Then there exists ( 𝜀 > 0 such) that 𝐵(𝑓 (𝑥), 𝜀) ⊆ 𝑉 .
By assumption there exists 𝛿 > 0 such that 𝑑𝑌 𝑓 (𝑥), 𝑓 (𝑦) ≤ 𝜀 for all 𝑦 ∈ 𝑋 with
𝑑𝑋 (𝑥, 𝑦) < 𝛿, that is, 𝑓 (𝑈 ) ⊆ 𝑉 if we let 𝑈 ∶= 𝐵(𝑥, 𝛿). As 𝑈 is a neighbourhood of 𝑥
it follows that 𝑓 is continuous. Let now 𝑓 be continuous and (𝑥𝑛 ) a sequence in 𝑋 con-
verging to 𝑥. If 𝜀 > 0 is given then there exists 𝛿 > 0 such that 𝑑𝑌 (𝑓 (𝑥), 𝑓 (𝑦)) < 𝜀 for all
𝑦 ∈ 𝑋 with 𝑑𝑋 (𝑥, 𝑦) < 𝛿. As 𝑥𝑛 → 𝑥 there exists 𝑛0 ∈ ℕ such that 𝑑𝑋 (𝑥, 𝑥𝑛 ) < 𝛿 for all
𝑛 ≥ 𝑛0 . Hence 𝑑𝑌 (𝑓 (𝑥), 𝑓 (𝑥𝑛 )) < 𝜀 for all 𝑛 ≥ 𝑛0 . As 𝜀 > 0 was arbitrary 𝑓 (𝑥𝑛 ) → 𝑓 (𝑥)
as 𝑛 → ∞. Assume now that (ii) does not hold. Then there exists 𝜀 > 0 such that for
each 𝑛 ∈ ℕ there exists 𝑥𝑛 ∈ 𝑋 with 𝑑𝑋 (𝑥, 𝑥𝑛 ) < 1∕𝑛 but 𝑑𝑌 (𝑓 (𝑥), 𝑓 (𝑥𝑛 )) ≥ 𝜀 for
all 𝑛 ∈ ℕ. Hence 𝑥𝑛 → 𝑥 in 𝑋 but 𝑓 (𝑥𝑛 ) ̸→ 𝑓 (𝑥) in 𝑌 , so (iii) does not hold. By
contrapositive (iii) implies (ii), completing the proof of the theorem.
(i) 𝑓 ∈ 𝐶(𝑋, 𝑌 );
(iv) For every 𝑥 ∈ 𝑋 and every neighbourhood 𝑉 ⊆ 𝑌 of 𝑓 (𝑥) there exists a neigh-
bourhood 𝑈 ⊆ 𝑋 of 𝑥 such that 𝑓 (𝑈 ) ⊆ 𝑉 ;
( )
(v) For every 𝑥 ∈ 𝑋 and every 𝜀 > 0 there exists 𝛿 > 0 such that 𝑑𝑌 𝑓 (𝑥), 𝑓 (𝑦) < 𝜀
for all 𝑦 ∈ 𝑋 with 𝑑𝑋 (𝑥, 𝑦) < 𝛿.
13
5.4 Definition (Distance to a Set) Let 𝐴 be a nonempty subset of 𝑋. We define the
distance between 𝑥 ∈ 𝑋 and 𝐴 by
5.5 Proposition For every nonempty set 𝐴 ⊆ 𝑋 the map 𝑋 → ℝ, 𝑥 → dist(𝑥, 𝐴), is
continuous.
Proof. By the properties of a metric 𝑑(𝑥, 𝑎) ≤ 𝑑(𝑥, 𝑦) + 𝑑(𝑦, 𝑎). By first taking
an infimum on the left hand side and then on the right hand side we get dist(𝑥, 𝐴) ≤
𝑑(𝑥, 𝑦) + dist(𝑦, 𝐴) and thus
5.6 Theorem If 𝑓 ∈ 𝐶(𝑋, 𝑌 ) and 𝑋 is compact then the image 𝑓 (𝑋) is compact in 𝑌 .
Proof. Suppose that (𝑈𝛼 ) is an open cover of 𝑓 (𝑋) then by continuity 𝑓 −1 [𝑈𝛼 ] are
open sets, and so (𝑓 −1 [𝑈𝛼 ]) is an open cover of 𝑋. By the compactness of 𝑋 it has a
finite sub-cover. Clearly the image of that finite sub-cover is a finite sub-cover of 𝑓 (𝑋)
by (𝑈𝛼 ). Hence 𝑓 (𝑋) is compact.
From the above theorem we can deduce an important property of real valued continuous
functions.
5.7 Theorem (Extreme value theorem) Suppose that 𝑋 is a compact metric space
and 𝑓 ∈ 𝐶(𝑋, ℝ). Then 𝑓 attains its maximum and minimum, that is, there exist
𝑥1 , 𝑥2 ∈ 𝑋 such that 𝑓 (𝑥1 ) = inf{𝑓 (𝑥) ∶ 𝑥 ∈ 𝑋} and 𝑓 (𝑥2 ) = sup{𝑓 (𝑥) ∶ 𝑥 ∈ 𝑋}.
Proof. By Theorem 5.6 the image of 𝑓 is compact, and so by the Heine-Borel theorem
(Theorem 4.5) closed and bounded. Hence the image 𝑓 (𝑋) = {𝑓 (𝑥) ∶ 𝑥 ∈ 𝑋} contain
its infimum and supremum, that is, 𝑥1 and 𝑥2 as required exist.
Continuous functions on compact sets have other nice properties. To discuss these
properties we introduce a stronger notion of continuity.
14
functions are uniformly continuous. For instance the function 𝑓 ∶ ℝ → ℝ given by
𝑓 (𝑥) = 𝑥2 is not uniformly continuous. To see this note that
for all 𝑥, 𝑦 ∈ ℝ. Hence, no matter how small |𝑥 − 𝑦| is, |𝑓 (𝑥) − 𝑓 (𝑦)| can be as big
as we like by choosing |𝑥 + 𝑦| large enough. However, the above also shows that 𝑓 is
uniformly continuous on any bounded set of ℝ.
We next show that continuous functions on compact metric spaces are automatically
uniformly continuous. The direct proof based on standard properties of continuous
functions is taken from [6].
𝛿 ∶= 𝑑𝑋 (𝑥0 , 𝑦0 ) ≤ 𝑑𝑋 (𝑥, 𝑦)
This is exactly what is required for uniform continuity. If 𝐴𝜀 = ∅, then (5.1) holds for
every 𝛿 > 0. As the arguments work for every choice of 𝜀 > 0 this proves the uniform
continuity of 𝑓 .
One could give an alternative proof of the above theorem using the covering property
of compact sets, or a contradiction proof based on the sequential compactness.
15
16
Chapter II
Banach Spaces
6 Normed Spaces
We consider a class of vector spaces with an additional topological structure. The
underlying field is always ℝ or ℂ. Most of the theory is developed simultaneously for
vector spaces over the two fields. Throughout, 𝕂 will be one of the two fields.
6.2 Proposition (Reversed triangle inequality) Let (𝐸, ‖⋅‖) be a normed space. Then,
for all 𝑥, 𝑦 ∈ 𝐸,
| |
‖𝑥 − 𝑦‖ ≥ |‖𝑥‖ − ‖𝑦‖|.
| |
Proof. By the triangle inequality ‖𝑥‖ = ‖𝑥 − 𝑦 + 𝑦‖ ≤ ‖𝑥 − 𝑦‖ + ‖𝑦‖, so ‖𝑥 − 𝑦‖ ≥
‖𝑥‖ − ‖𝑦‖. Interchanging the roles of 𝑥 and 𝑦 and applying (ii) we get ‖𝑥 − 𝑦‖ =
| − 1|‖(−1)(𝑥 − 𝑦)‖ = ‖𝑦 − 𝑥‖ ≥ ‖𝑦‖ − ‖𝑥‖. Combining the two inequalities, the
assertion of the proposition follows.
17
6.3 Lemma Let (𝐸, ‖⋅‖) be a normed space and define
𝑑(𝑥, 𝑦) ∶= ‖𝑥 − 𝑦‖
We will always equip a normed space with the topology of 𝐸 generated by the metric
induced by the norm. Hence it makes sense to talk about continuity of functions. It
turns out the that topology is compatible with the vector space structure as the following
theorem shows.
6.4 Theorem Given a normed space (𝐸, ‖⋅‖), the following maps are continuous (with
respect to the product topologies).
6.5 Definition (Banach space) A normed space which is complete with respect to the
metric induced by the norm is called a Banach space.
6.6 Example The simplest example of a Banach space is ℝ𝑁 or ℂ𝑁 with the Euclidean
norm.
18
We give a characterisation of Banach spaces in terms of properties of series. We recall
the following definition.
∑∞
6.7 Definition (absolute convergence) A series 𝑘=0 𝑎𝑘 in 𝐸 is called absolutely con-
∑∞
vergent if 𝑘=0 ‖𝑎𝑘 ‖ converges.
6.8 Theorem A normed space 𝐸 is complete if and only every absolutely convergent
series in 𝐸 converges.
∑∞
Proof. Suppose that 𝐸 is complete. Let 𝑘=0 𝑎𝑘 an absolutely convergent series in 𝐸,
∑∞
that is, 𝑘=0 ‖𝑎𝑘 ‖ converges. By the Cauchy criterion for the convergence of a series
in ℝ, for every 𝜀 > 0 there exists 𝑛0 ∈ ℕ such that
∑
𝑛
‖𝑎𝑘 ‖ < 𝜀
𝑘=𝑚+1
∑𝑛
for all 𝑛 > 𝑚 > 𝑛0 . (This simply means that the sequence of partial sums 𝑘=0
‖𝑎𝑘 ‖,
𝑛 ∈ ℕ is a Cauchy sequence.) Therefore, by the triangle inequality
‖∑ ∑ ‖ ‖∑ ‖ ∑
𝑛 𝑚 𝑛 𝑛
‖ 𝑎𝑘 − 𝑎𝑘 ‖ = ‖ 𝑎𝑘 ‖ ≤ ‖𝑎𝑘 ‖ < 𝜀
‖ ‖ ‖ ‖
𝑘=0 𝑘=0 𝑘=𝑚+1 𝑘=𝑚+1
∑𝑛
for all 𝑛 > 𝑚 > 𝑛0 . Hence the sequence of partial sums 𝑘=0 𝑎𝑘 , 𝑛 ∈ ℕ is a Cauchy
sequence in 𝐸. Since 𝐸 is complete,
∑
∞
∑
𝑛
𝑎𝑘 = lim 𝑎𝑘
𝑛→∞
𝑘=0 𝑘=0
19
7 Examples of Banach Spaces
In this section we give examples of Banach spaces. They will be used throughout the
course. We start by elementary inequalities and a family of norms in 𝕂𝑁 . They serve
as a model for more general spaces of sequences or functions.
⎧(∑𝑁 )1∕𝑝
⎪ |𝑥𝑖 |𝑝 if 1 ≤ 𝑝 < ∞,
|𝑥|𝑝 ∶= ⎨ 𝑖=1 (7.1)
⎪ max |𝑥𝑖 | if 𝑝 = ∞.
⎩ 𝑖=1,…,𝑁
At this stage we do not know whether |⋅|𝑝 is a norm. We now prove that the 𝑝-norms are
norms, and derive some relationships between them. First we need Young’s inequality.
The relationship (7.2) is rather important and appears very often. We say that 𝑝′ is the
exponent dual to 𝑝. If 𝑝 = 1 we set 𝑝′ ∶= ∞ and if 𝑝 = ∞ we set 𝑝′ ∶= 1.
From Young’s inequality we get Hölder’s inequality.
7.2 Proposition Let 1 ≤ 𝑝 ≤ ∞ and 𝑝′ the exponent dual to 𝑝. Then for all 𝑥, 𝑦 ∈ 𝕂𝑁
∑
𝑁
|𝑥𝑖 ||𝑦𝑖 | ≤ |𝑥|𝑝 |𝑦|𝑝′ .
𝑖=1
20
Proof. If 𝑥 = 0 or 𝑦 = 0 then the inequality is obvious. Also if 𝑝 = 1 or 𝑝 = ∞ then
the inequality is also rather obvious. Hence assume that 𝑥, 𝑦 ≠ 0 and 1 < 𝑝 < ∞. By
Young’s inequality (Lemma 7.1) we have
𝑁 ( ( ) ( ) ′)
1 ∑𝑁
∑𝑁
|𝑥𝑖 | |𝑦𝑖 | ∑ 1 |𝑥𝑖 | 𝑝 1 |𝑦𝑖 | 𝑝
|𝑥𝑖 ||𝑦𝑖 | = ≤ + ′
|𝑥|𝑝 |𝑦|𝑝′ 𝑖=1 𝑖=1
|𝑥|𝑝 |𝑦|𝑝′ 𝑖=1
𝑝 |𝑥|𝑝 𝑝 |𝑦|𝑝′
𝑁 ( ) 𝑁 ( )′
1 ∑ |𝑥𝑖 | 𝑝 1 ∑ |𝑦𝑖 | 𝑝
= + ′
𝑝 𝑖=1 |𝑥|𝑝 𝑝 𝑖=1 |𝑦|𝑝′
1 1 ∑ 1 1 ∑
𝑁 𝑁
′
= 𝑝 |𝑥𝑖 | + ′ 𝑝′
𝑝
|𝑦𝑖 |𝑝
𝑝 |𝑥|𝑝 𝑖=1 𝑝 |𝑦| ′ 𝑖=1
𝑝
𝑝 ′
1 |𝑥|𝑝 1 |𝑦|𝑝′
𝑝
1 1
= 𝑝 + ′ = + ′ = 1,
𝑝 |𝑥|𝑝 𝑝 |𝑦| ′
𝑝′
𝑝 𝑝
𝑝
(∑
𝑁 )1∕𝑝 (∑
𝑁 )1∕𝑝
|𝛼𝑥|𝑝 = |𝛼𝑥𝑖 | 𝑝
= |𝛼|𝑝 |𝑥𝑖 |𝑝 = |𝛼||𝛼𝑥|𝑝 .
𝑖=1 𝑖=1
Thus it remains to prove the triangle inequality. For 𝑥, 𝑦 ∈ 𝕂𝑁 we have, using Hölder’s
inequality (Proposition 7.2), that
∑
𝑁
∑
𝑁
|𝑥 + 𝑦|𝑝𝑝 = |𝑥𝑖 + 𝑦𝑖 | =𝑝
|𝑥𝑖 + 𝑦𝑖 ||𝑥𝑖 + 𝑦𝑖 |𝑝−1
𝑖=1 𝑖=1
∑𝑁
∑
𝑁
≤ |𝑥𝑖 ||𝑥𝑖 + 𝑦𝑖 | 𝑝−1
+ |𝑦𝑖 ||𝑥𝑖 + 𝑦𝑖 |𝑝−1
𝑖=1 𝑖=1
(∑
𝑁 )1∕𝑝′ (∑
𝑁 )1∕𝑝′
(𝑝−1)𝑝′ (𝑝−1)𝑝′
≤ |𝑥|𝑝 |𝑥𝑖 + 𝑦𝑖 | + |𝑦|𝑝 |𝑥𝑖 + 𝑦𝑖 |
𝑖=1 𝑖=1
( )(∑
𝑁
(𝑝−1)𝑝′
)1∕𝑝′
= |𝑥|𝑝 + |𝑦|𝑝 |𝑥𝑖 + 𝑦𝑖 | .
𝑖=1
21
Now, observe that
( )−1 𝑝
′ 1
𝑝 = 1− = ,
𝑝 𝑝−1
so we get from the above that
( )(∑
𝑁 )(𝑝−1)∕𝑝 ( )
|𝑥 + 𝑦|𝑝𝑝 ≤ |𝑥|𝑝 + |𝑦|𝑝 |𝑥𝑖 + 𝑦𝑖 |𝑝 = |𝑥|𝑝 + |𝑦|𝑝 |𝑥 + 𝑦|𝑝−1
𝑝
.
𝑖=1
Hence if 𝑥 + 𝑦 ≠ 0 we get the triangle inequality |𝑥 + 𝑦|𝑝 ≤ |𝑥|𝑝 + |𝑦|𝑝 . The inequality
is obvious if 𝑥 + 𝑦 = 0, so |⋅|𝑝 is a norm on 𝕂𝑁 .
Next we show the first inequality in (7.3). First let 𝑝 < 𝑞 = ∞. If 𝑥 ∈ 𝕂𝑁 we pick
( )1∕𝑝
the component 𝑥𝑗 of 𝑥 such that |𝑥𝑗 | = |𝑥|∞ . Hence |𝑥|∞ = |𝑥𝑗 | = |𝑥𝑗 |𝑝 ≤ |𝑥|𝑝 ,
proving the first inequality in case 𝑞 = ∞. Assume now that 1 ≤ 𝑝 ≤ 𝑞 < ∞. If 𝑥 ≠ 0
then |𝑥𝑖 |∕|𝑥|𝑝 ≤ 1 Hence, as 1 ≤ 𝑝 ≤ 𝑞 < ∞ we have
( |𝑥 | )𝑞 ( |𝑥 | )𝑝
𝑖 𝑖
≤
|𝑥|𝑝 |𝑥|𝑝
for all 𝑥 ∈ 𝕂𝑁 ⧵ {0}. Therefore,
𝑁 (
|𝑥𝑖 | )𝑞
𝑁 (
|𝑥|𝑞𝑞 ∑ ∑ |𝑥𝑖 | )𝑝 |𝑥|𝑝𝑝
= ≤ = =1
|𝑥|𝑞𝑝 𝑖=1
|𝑥|𝑝 𝑖=1
|𝑥|𝑝 |𝑥|𝑝𝑝
for all 𝑥 ∈ 𝕂𝑁 ⧵ {0}. Hence |𝑥|𝑞 ≤ |𝑥|𝑝 for all 𝑥 ≠ 0. For 𝑥 = 0 the inequality is
trivial. To prove the second inequality in (7.3) assume that 1 ≤ 𝑝 < 𝑞 < ∞. We define
𝑠 ∶= 𝑞∕𝑝. The corresponding dual exponent 𝑠′ is given by
𝑠 𝑞
𝑠′ = = .
𝑠−1 𝑞−𝑝
Applying Hölder’s inequality we get
∑
𝑁 (∑
𝑁 )1∕𝑠 (∑
𝑁 )1∕𝑠′ 𝑞−𝑝
(∑
𝑁 )𝑝∕𝑞 𝑞−𝑝
𝑠′
|𝑥|𝑝𝑝 = |𝑥𝑖 | ⋅ 1 ≤
𝑝
|𝑥𝑖 | 𝑝𝑠
1 =𝑁 𝑞 |𝑥𝑖 |𝑞 =𝑁 𝑞 |𝑥|𝑝𝑞
𝑖=1 𝑖=1 𝑖=1 𝑖=1
for all 𝑥 ∈ 𝕂𝑁 , from which the second inequality in (7.3) follows. If 1 ≤ 𝑝 < 𝑞 = ∞
and 𝑥 ∈ 𝕂𝑁 is given we pick 𝑥𝑗 such that |𝑥𝑗 | = |𝑥|∞ . Then
(∑
𝑁 )1∕𝑝 (∑
𝑁 )1∕𝑝
|𝑥|𝑝 = |𝑥𝑖 | 𝑝
≤ |𝑥𝑗 |𝑝 = 𝑁 1∕𝑝 |𝑥𝑗 | = 𝑁 1∕𝑝 |𝑥|∞ ,
𝑖=1 𝑖=1
22
7.2 Spaces of Sequences
Here we discuss spaces of sequences. As most of you have seen this in Metric Spaces
the exposition will be rather brief.
Denote by the space of all sequences in 𝕂, that is, the space of all functions from
ℕ into 𝕂. We denote its elements by 𝑥 = (𝑥0 , 𝑥1 , 𝑥2 , … ) = (𝑥𝑖 ). We define vector space
operations “component” wise:
⎧(∑∞ )1∕𝑝
⎪ |𝑥𝑖 |𝑝 if 1 ≤ 𝑝 < ∞,
|(𝑥𝑖 )|𝑝 ∶= ⎨ 𝑖=1 (7.4)
⎪sup |𝑥 | if 𝑝 = ∞.
⎩ 𝑖∈ℕ 𝑖
These 𝑝-norms are not finite for all sequences. We define some subspaces of in the
following way:
{ }
• 𝓁𝑝 ∶= 𝓁𝑝 (𝕂) ∶= 𝑥 ∈ ∶ |𝑥|𝑝 < ∞ (1 ≤ 𝑝 ≤ ∞);
{ }
• 𝑐0 ∶= 𝑐0 (𝕂) ∶= (𝑥𝑖 ) ∈ ∶ lim |𝑥𝑖 | = 0 .
𝑖→∞
The 𝑝-norms for sequences have similar properties as the 𝑝-norms in 𝕂𝑁 . In fact most
properties follow from the finite version given in Section 7.1.
∑
∞
|𝑥𝑖 ||𝑦𝑖 | ≤ |𝑥|𝑝 |𝑦|𝑝′ ,
𝑖=1
7.5 Theorem Let 1 ≤ 𝑝 ≤ ∞. Then (𝓁𝑝 , |⋅|𝑝 ) and (𝑐0 , , |⋅|∞ ) are Banach spaces. More-
over, if 1 < 𝑝 < 𝑞 < ∞ then
𝓁1 ⊊ 𝓁𝑝 ⊊ 𝓁𝑞 ⊊ 𝑐0 ⊊ 𝓁∞ .
23
. If |(𝑥𝑖 )|𝑝 < ∞ for some (𝑥𝑖 ) ∈ and 𝑝 < ∞ then we must have |𝑥𝑖 | → 0. Hence
𝓁𝑝 ⊂ 𝑐0 for all 1 ≤ 𝑝 < ∞. Clearly 𝑐0 ⊊ 𝓁∞ . If 1 ≤ 𝑝 < 𝑞 < ∞ then by Theorem 7.3
(∑
𝑛 )1∕𝑞 (∑
𝑛 )1∕𝑝
sup |𝑥𝑖 | ≤ |𝑥𝑖 |
𝑞
≤ |𝑥𝑖 |𝑝 .
𝑖=1,…,𝑛 𝑖=1 𝑖=1
Passing to the limit on the right hand side and then taking the supremum on the left
hand side we get
|(𝑥𝑖 )|∞ ≤ |(𝑥𝑖 )|𝑞 ≤ |(𝑥𝑖 )|𝑝 ,
proving the inclusions and the inequalities. To show that the inclusions are proper we
use the harmonic series
∑∞
1
= ∞.
𝑖=1
𝑖
Clearly (1∕𝑖) ∈ 𝑐0 but not in 𝓁1 . Similarly, (1∕𝑖1∕𝑝 ) ∈ 𝓁𝑞 for 𝑞 > 𝑝 but not in 𝓁𝑝 .
We finally prove completeness. Suppose that (𝑥𝑛 ) is a Cauchy sequence in 𝓁𝑝 . Then
by definition of the 𝑝-norm
for all 𝑖, 𝑚, 𝑛 ∈ ℕ. It follows that (𝑥𝑖𝑛 ) is a Cauchy sequence in 𝕂 for every 𝑖 ∈ ℕ. Since
𝕂 is complete
𝑥𝑖 ∶= lim 𝑥𝑖𝑛 (7.5)
𝑛→∞
exists for all 𝑖 ∈ ℕ. We set 𝑥 ∶= (𝑥𝑖 ). We need to show that 𝑥𝑛 → 𝑥 in 𝓁𝑝 , that is, with
respect to the 𝓁𝑝 -norm. Let 𝜀 > 0 be given. By assumption there exists 𝑛0 ∈ ℕ such
that for all 𝑛, 𝑚 > 𝑛0
(∑
𝑁 )1∕𝑝
𝜀
|𝑥𝑖𝑛 − 𝑥𝑖𝑚 |𝑝 ≤ |𝑥𝑛 − 𝑥𝑚 |𝑝 <
𝑖=1
2
if 1 ≤ 𝑝 < ∞ and
𝜀
max |𝑥𝑖𝑛 − 𝑥𝑖𝑚 | ≤ |𝑥𝑛 − 𝑥𝑚 |∞ <
𝑖=1…,𝑁 2
if 𝑝 = ∞. For fixed 𝑁 ∈ ℕ we can let 𝑚 → ∞, so by (7.5) and the continuity of the
absolute value, for all 𝑁 ∈ ℕ and 𝑛 > 𝑛0
(∑
𝑁 )1∕𝑝
𝜀
|𝑥𝑖𝑛 − 𝑥𝑖 |𝑝 ≤
𝑖=1
2
if 1 ≤ 𝑝 < ∞ and
𝜀
max |𝑥𝑖𝑛 − 𝑥𝑖𝑚 | ≤ |𝑥𝑛 − 𝑥|∞ ≤
𝑖=1…,𝑁 2
if 𝑝 = ∞. Letting 𝑁 → ∞ we finally get
𝜀
|𝑥𝑛 − 𝑥|𝑝 ≤ <𝜀
2
for all 𝑛 > 𝑛0 . Since the above works for every 𝜀 > 0 it follows that 𝑥𝑛 − 𝑥 → 0
in 𝓁𝑝 as 𝑛 → ∞. Finally, as 𝓁𝑝 is a vector space and 𝑥𝑛0 , 𝑥𝑛0 − 𝑥 ∈ 𝓁𝑝 we have
24
𝑥 = 𝑥𝑛0 − (𝑥𝑛0 − 𝑥) ∈ 𝓁𝑝 . Hence 𝑥𝑛 → 𝑥 in 𝓁𝑝 , showing that 𝓁𝑝 is complete for
1 ≤ 𝑝 ≤ ∞. To show that 𝑐0 is complete we need to show that 𝑥 ∈ 𝑐0 if 𝑥𝑛 ∈ 𝑐0 for all
𝑛 ∈ ℕ. We know that 𝑥𝑛 → 𝑥 in 𝓁∞ . Hence, given 𝜀 > 0 there exists 𝑛0 ∈ ℕ such that
|𝑥𝑛 − 𝑥|∞ < 𝜀∕2 for all 𝑛 ≥ 𝑛0 . Therefore
𝜀
|𝑥𝑖 | ≤ |𝑥𝑖 − 𝑥𝑖𝑛0 | + |𝑥𝑖𝑛0 | ≤ |𝑥 − 𝑥𝑛0 |∞ + |𝑥𝑖𝑛0 | < + |𝑥𝑖𝑛0 |
2
for all 𝑖 ∈ ℕ. Since 𝑥𝑛0 ∈ 𝑐0 there exists 𝑖0 ∈ ℕ such that |𝑥𝑖𝑛0 | < 𝜀∕2 for all 𝑖 > 𝑖0 .
Hence |𝑥𝑖 | ≤ 𝜀∕2 + 𝜀∕2 = 𝜀 for all 𝑖 > 𝑖0 , so 𝑥 ∈ 𝑐0 as claimed. This completes the
proof of completeness of 𝓁𝑝 and 𝑐0 .
7.6 Remark The proof of completeness in many cases follows similar steps as the one
above.
(1) Take a an arbitrary Cauchy sequence (𝑥𝑛 ) in the normed space 𝐸 and show that (𝑥𝑛 )
converges not in the norm of 𝐸, but in some weaker sense to some 𝑥. (In the above
proof it is, “component-wise” convergence, that is 𝑥𝑖𝑛 converges for each 𝑖 ∈ ℕ to
some 𝑥𝑖 .);
(2) Show that ‖𝑥𝑛 − 𝑥‖𝐸 → 0;
(3) Show that 𝑥 ∈ 𝐸 by using that 𝐸 is a vector space.
7.8 Theorem Let 1 ≤ 𝑝 ≤ ∞. Then (𝐿𝑝 (𝑋), ‖⋅‖𝑝 ) is a Banach space. If 𝑋 has finite
measure then
𝐿∞ (𝑋) ⊂ 𝐿𝑞 (𝑋) ⊂ 𝐿𝑝 (𝑋) ⊂ 𝐿1 (𝑋)
if 1 < 𝑝 < 𝑞 < ∞. If 𝑋 has infinite measure there are no inclusions between 𝐿𝑝 (𝑋)
and 𝐿𝑞 (𝑋) for 𝑝 ≠ 𝑞.
25
7.4 Spaces of Bounded and Continuous Functions
Suppose that 𝑋 is a set and 𝐸 = (𝐸, ‖⋅‖) a normed space. For a function 𝑢 ∶ 𝑋 → 𝐸
we let
‖𝑢‖∞ ∶= sup ‖𝑢(𝑥)‖
𝑥∈𝑋
7.9 Theorem If 𝑋 is a set and 𝐸 a Banach space, then 𝐵(𝑋, 𝐸) is a Banach space
with the supremum norm.
Proof. Let (𝑢𝑛 ) be a Cauchy sequence in 𝐵(𝑋, 𝐸). As
exists for all 𝑥 ∈ 𝑋. We need to show that 𝑢𝑛 → 𝑢 in 𝐵(𝑋, 𝐸), that is, with respect to
the supremum norm. Let 𝜀 > 0 be given. Then by assumption there exists 𝑛0 ∈ ℕ such
that ‖𝑢𝑛 − 𝑢𝑚 ‖∞ < 𝜀∕2 for all 𝑚, 𝑛 > 𝑛0 . Using (7.7) we get
𝜀
‖𝑢𝑛 (𝑥) − 𝑢𝑚 (𝑥)‖ <
2
for all 𝑚, 𝑛 > 𝑛0 and 𝑥 ∈ 𝑋. For fixed 𝑥 ∈ 𝑋 we can let 𝑚 → ∞, so by the continuity
of the norm
𝜀
‖𝑢𝑛 (𝑥) − 𝑢(𝑥)‖ <
2
for all 𝑥 ∈ 𝑋 and all 𝑛 > 𝑛0 . Hence ‖𝑢𝑛 −𝑢‖∞ ≤ 𝜀∕2 < 𝜀 for all 𝑛 > 𝑛0 . Since the above
works for every 𝜀 > 0 it follows that 𝑢𝑛 − 𝑢 → 0 in 𝐸 as 𝑛 → ∞. Finally, as 𝐵(𝑋, 𝐸)
is a vector space and 𝑢𝑛0 , 𝑢𝑛0 − 𝑢 ∈ 𝐵(𝑋, 𝐸) we have 𝑢 = 𝑢𝑛0 − (𝑢𝑛0 − 𝑢) ∈ 𝐵(𝑋, 𝐸).
Hence 𝑢𝑛 → 𝑢 in 𝐵(𝑋, 𝐸), showing that 𝐵(𝑋, 𝐸) is complete.
If 𝑋 is a metric space we denote the vector space of all continuous functions by 𝐶(𝑋, 𝐸).
This space does not carry a topology or norm in general. However,
becomes a normed space with norm ‖⋅‖∞ . Note that if 𝑋 is compact then 𝐶(𝑋, 𝐸) =
𝐵𝐶(𝑋, 𝐸). The space 𝐵𝐶(𝑋, 𝐸) turns out to be a Banach space if 𝐸 is a Banach
space. Note that the completeness of 𝐵𝐶(𝑋, 𝐸) is equivalent to the fact that the uniform
limit of continuous functions is continuous. Hence the language of functional analysis
provides a way to rephrase standard facts from analysis in a concise and unified way.
26
7.10 Theorem If 𝑋 is a metric space and 𝐸 a Banach space, then 𝐵𝐶(𝑋, 𝐸) is a
Banach space with the supremum norm.
‖𝑢(𝑥) − 𝑢(𝑥0 )‖𝐸 ≤ ‖𝑢(𝑥) − 𝑢𝑛 (𝑥)‖𝐸 + ‖𝑢𝑛 (𝑥) − 𝑢𝑛 (𝑥0 )‖𝐸 + ‖𝑢𝑛 (𝑥0 ) − 𝑢(𝑥0 )‖𝐸 (7.8)
for all 𝑥 ∈ 𝑋 and 𝑛 ∈ ℕ. Fix now 𝜀 > 0 arbitrary. Since 𝑢𝑛 → 𝑢 in 𝐵(𝑋, 𝐸) there
exists 𝑛0 ∈ ℕ such that ‖𝑢𝑛 (𝑥) − 𝑢(𝑥)‖𝐸 < 𝜀∕4 for all 𝑛 > 𝑛0 and 𝑥 ∈ 𝑋. Hence (7.8)
implies that
𝜀
‖𝑢(𝑥) − 𝑢(𝑥0 )‖𝐸 < + ‖𝑢𝑛 (𝑥) − 𝑢𝑛 (𝑥0 )‖𝐸 . (7.9)
2
Since 𝑢𝑛0 +1 is continuous at 𝑥0 there exists 𝛿 > 0 such that ‖𝑢𝑛0 +1 (𝑥)−𝑢𝑛0 +1 (𝑥0 )‖𝐸 < 𝜀∕2
for all 𝑥 ∈ 𝑋 with 𝑑(𝑥, 𝑥0 ) < 𝛿 and 𝑛 > 𝑛0 . Using (7.9) we get ‖𝑢(𝑥) − 𝑢(𝑥0 )‖𝐸 <
𝜀∕2+𝜀∕2 if 𝑑(𝑥, 𝑥0 ) < 𝛿 and so 𝑢 is continuous at 𝑥0 . As 𝑥0 was arbitrary, 𝑢 ∈ 𝐶(𝑋, 𝐸)
as claimed.
Note that the above is a functional analytic reformulation of the fact that a uniformly
convergent sequence of bounded functions is bounded, and similarly that a uniformly
convergent sequence of continuous functions is continuous.
In the following theorem we collect the main properties of continuous and bounded
linear operators. In particular we show that a linear operator is bounded if and only if
it is continuous.
27
8.3 Theorem For 𝑇 ∈ Hom(𝐸, 𝐹 ) the following statements are equivalent:
(ii) 𝑇 ∈ (𝐸, 𝐹 );
(iii) 𝑇 is continuous at 𝑥 = 0;
(iv) 𝑇 is bounded;
(v) There exists 𝛼 > 0 such that ‖𝑇 𝑥‖𝐹 ≤ 𝛼‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸.
Proof. The implications (i)⇒(ii) and (ii)⇒(iii) are obvious. Suppose now that 𝑇 is
continuous at 𝑥 = 0 and that 𝑈 is an arbitrary bounded subset of 𝐸. As 𝑇 is continuous
at 𝑥 = 0 there exists 𝛿 > 0 such that ‖𝑇 𝑥‖𝐹 ≤ 1 whenever ‖𝑥‖𝐸 ≤ 𝛿. Since 𝑈 is
bounded 𝑀 ∶= sup𝑥∈𝑈 ‖𝑥‖𝐸 < ∞, so for every 𝑥 ∈ 𝑈
‖𝛿 ‖ 𝛿
‖ 𝑥‖ = ‖𝑥‖𝐸 ≤ 𝛿.
‖ 𝑀 ‖𝐸 𝑀
Hence by the linearity of 𝑇 and the choice of 𝛿
( )
𝛿 ‖ 𝛿 ‖
‖𝑇 𝑥‖𝐹 = ‖𝑇 𝑥 ‖ ≤ 1,
𝑀 ‖ 𝑀 ‖𝐹
showing that
𝑀
‖𝑇 𝑥‖𝐹 ≤
𝛿
for all 𝑥 ∈ 𝑈 . Therefore, the image of 𝑈 under 𝑇 is bounded, showing that (iii) implies
(iv). Suppose now that 𝑇 is bounded. Then there exists 𝛼 > 0 such that ‖𝑇 𝑥‖𝐹 ≤ 𝛼
whenever ‖𝑥‖𝐸 ≤ 1. Hence, using the linearity of 𝑇
( )
1 ‖ 𝑥 ‖
‖𝑇 𝑥‖𝐹 = ‖𝑇 ‖ ≤𝛼
‖𝑥‖𝐸 ‖ ‖𝑥‖𝐸 ‖𝐹
for all 𝑥 ∈ 𝐸 with 𝑥 ≠ 0. Since 𝑇 0 = 0 it follows that ‖𝑇 𝑥‖𝐹 ≤ 𝛼‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸.
Hence (iv) implies (v). Suppose now that there exists 𝛼 > 0 such that ‖𝑇 𝑥‖𝐹 ≤ 𝛼‖𝑥‖𝐸
for all 𝑥 ∈ 𝐸. Then by the linearity of 𝑇
showing the uniform continuity of 𝑇 . Hence (v) implies (i), completing the proof of
the theorem.
28
8.5 Remark We could define ‖𝑇 ‖(𝐸,𝐹 ) for all 𝑇 ∈ Hom(𝐸, 𝐹 ), but Theorem 8.3
shows that 𝑇 ∈ (𝐸, 𝐹 ) if and only if ‖𝑇 ‖(𝐸,𝐹 ) < ∞.
Before proving that the operator norm is in fact a norm, we first give other characteri-
sations.
8.6 Proposition Suppose that 𝑇 ∈ (𝐸, 𝐹 ). Then ‖𝑇 𝑥‖𝐹 ≤ ‖𝑇 ‖(𝐸,𝐹 ) ‖𝑥‖𝐸 for all
𝑥 ∈ 𝐸. Moreover,
‖𝑇 𝑥‖𝐹
‖𝑇 ‖(𝐸,𝐹 ) = sup = sup ‖𝑇 𝑥‖𝐹 = sup ‖𝑇 𝑥‖𝐹 = sup ‖𝑇 𝑥‖𝐹 . (8.1)
𝑥∈𝐸⧵{0} ‖𝑥‖𝐸 ‖𝑥‖𝐸 =1 ‖𝑥‖𝐸 <1 ‖𝑥‖𝐸 ≤1
Proof. Fix 𝑇 ∈ (𝐸, 𝐹 ) and set 𝐴 ∶= {𝛼 > 0 ∶ ‖𝑇 𝑥‖𝐹 ≤ 𝛼‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸}. By
definition ‖𝑇 ‖(𝐸,𝐹 ) = inf 𝐴. If 𝛼 ∈ 𝐴, then ‖𝑇 𝑥‖𝐹 ≤ 𝛼‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸. Hence,
for every 𝑥 ∈ 𝐸 we have ‖𝑇 𝑥‖𝐹 ≤ (inf 𝐴)‖𝑥‖𝐸 , proving the first claim. Set now
‖𝑇 𝑥‖𝐹
𝜆 ∶= sup .
𝑥∈𝐸⧵{0} ‖𝑥‖𝐸
Then, ‖𝑇 𝑥‖𝐹 ≤ 𝜆‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸, and so 𝜆 ≥ inf 𝐴 = ‖𝑇 ‖(𝐸,𝐹 ) . By the above we
have ‖𝑇 𝑥‖𝐹 ≤ ‖𝑇 ‖(𝐸,𝐹 ) ‖𝑥‖𝐸 and thus
‖𝑇 𝑥‖𝐹
≤ ‖𝑇 ‖(𝐸,𝐹 )
‖𝑥‖𝐸
for all 𝑥 ∈ 𝐸 ⧵ {0}, implying that 𝜆 ≤ ‖𝑇 ‖(𝐸,𝐹 ) . Combining the inequalities 𝜆 =
‖𝑇 ‖(𝐸,𝐹 ) , proving the first equality in (8.1). Now by the linearity of 𝑇
‖𝑇 𝑥‖𝐹 ‖ 𝑥 ‖
sup = sup ‖𝑇 ‖ = sup ‖𝑇 𝑥‖𝐹 ,
𝑥∈𝐸⧵{0} ‖𝑥‖𝐸 𝑥∈𝐸⧵{0} ‖ ‖𝑥‖𝐸 ‖𝐹 ‖𝑥‖𝐸 =1
proving the second equality in (8.1). To prove the third equality note that
‖𝑇 𝑥‖𝐹 ‖𝑇 𝑥‖𝐹
𝛽 ∶= sup ‖𝑇 𝑥‖𝐹 ≤ sup ≤ sup = 𝜆.
‖𝑥‖𝐸 <1 ‖𝑥‖𝐸 <1 ‖𝑥‖𝐸 𝑥∈𝐸⧵{0} ‖𝑥‖𝐸
and thus ‖𝑇 𝑥‖𝐹 ≤ 𝛽(‖𝑥‖𝐸 + 𝜀) for all 𝜀 > 0 and 𝑥 ∈ 𝐸. Hence ‖𝑇 𝑥‖𝐹 ≤ 𝛽‖𝑥‖𝐸 for
all 𝑥 ∈ 𝐸, implying that 𝛽 ≥ 𝜆. Combining the inequalities 𝛽 = 𝜆, which is the third
inequality in (8.1). For the last equality note that ‖𝑇 𝑥‖𝐹 ≤ ‖𝑇 ‖(𝐸,𝐹 ) ‖𝑥‖𝐸 ≤ ‖𝑇 ‖(𝐸,𝐹 )
whenever ‖𝑥‖𝐸 ≤ 1. Hence
29
( )
8.7 Proposition The space (𝐸, 𝐹 ), ‖⋅‖(𝐸,𝐹 ) is a normed space.
Proof. By definition ‖𝑇 ‖(𝐸,𝐹 ) ≥ 0 for all 𝑇 ∈ (𝐸, 𝐹 ). Let now ‖𝑇 ‖(𝐸,𝐹 ) = 0. Then
by (8.1) we have
‖𝑇 𝑥‖𝐹
sup = 0,
𝑥∈𝐸⧵{0} ‖𝑥‖𝐸
From now on we will always assume that (𝐸, 𝐹 ) is equipped with the operator norm.
Using the steps outlined in Remark 7.6 we prove that (𝐸, 𝐹 ) is complete if 𝐹 is com-
plete.
8.8 Theorem If 𝐹 is a Banach space, then (𝐸, 𝐹 ) is a Banach space with respect to
the operator norm.
Proof. To simplify notation we let ‖𝑇 ‖ ∶= ‖𝑇 ‖(𝐸,𝐹 ) for all 𝑇 ∈ (𝐸, 𝐹 ). Sup-
pose that 𝐹 is a Banach space, and that (𝑇𝑛 ) is a Cauchy sequence in (𝐸, 𝐹 ). By
Proposition 8.6 we have
𝑇 𝑥 ∶= lim 𝑇𝑛 𝑥
𝑛→∞
30
for all 𝑛, 𝑚 ≥ 𝑛0 and all 𝑥 ∈ 𝐸. Letting 𝑚 → ∞ and using the continuity of the norm
we see that
‖𝑇𝑛 𝑥 − 𝑇 𝑥‖𝐹 ≤ 𝜀‖𝑥‖𝐸
for all 𝑥 ∈ 𝐸 and 𝑛 ≥ 𝑛0 . By definition of the operator norm ‖𝑇𝑛 −𝑇 ‖ ≤ 𝜀 for all 𝑛 ≥ 𝑛0 .
In particular, 𝑇𝑛 − 𝑇 ∈ (𝐸, 𝐹 ) for all 𝑛 ≥ 𝑛0 . Since 𝜀 > 0 was arbitrary, ‖𝑇𝑛 − 𝑇 ‖ → 0
in (𝐸, 𝐹 ). Finally, since (𝐸, 𝐹 ) is a vector space and 𝑇𝑛0 , 𝑇𝑛0 − 𝑇 ∈ (𝐸, 𝐹 ), we
have 𝑇 = 𝑇𝑛0 − (𝑇𝑛0 − 𝑇 ) ∈ (𝐸, 𝐹 ), completing the proof of the theorem.
9 Equivalent Norms
Depending on the particular problem we look at, it may be convenient to work with
different norms. Some norms generate the same topology as the original norm, others
may generate a different topology. Here are some definitions.
9.1 Definition (Equivalent norms) Suppose that 𝐸 is a vector space, and that ‖⋅‖1 and
‖⋅‖2 are norms on 𝐸.
• We say that ‖⋅‖1 is stronger than ‖⋅‖2 if there exists a constant 𝐶 > 0 such that
‖𝑥‖2 ≤ 𝐶‖𝑥‖1
for all 𝑥 ∈ 𝐸. In that case we also say that ‖⋅‖2 is weaker than ‖⋅‖1 .
• We say that the norms ‖⋅‖1 and ‖⋅‖2 are equivalent if there exist two constants
𝑐, 𝐶 > 0 such that
𝑐‖𝑥‖1 ≤ ‖𝑥‖2 ≤ 𝐶‖𝑥‖1
for all 𝑥 ∈ 𝐸.
9.2 Examples (a) Let |⋅|𝑝 denote the 𝑝-norms on 𝕂𝑁 , where 1 ≤ 𝑝 ≤ ∞ as defined in
Section 7.2. We proved in Theorem 7.3 that
𝑞−𝑝
|𝑥|𝑞 ≤ |𝑥|𝑝 ≤ 𝑁 𝑝𝑞 |𝑥|𝑞
31
(d) Consider the two normed spaces 𝐸1 ∶= (𝐸, ‖⋅‖1 ) and 𝐸2 ∶= (𝐸, ‖⋅‖2 ). Clearly
𝐸1 = 𝐸2 = 𝐸 as sets, but not as normed (or metric) spaces. By Theorem 8.3 it is
obvious that ‖⋅‖1 is stronger than ‖⋅‖2 if and only if the linear map 𝑖(𝑥) ∶= 𝑥 is a
bounded linear operator 𝑖 ∈ (𝐸1 , 𝐸2 ). If the two norms are equivalent then also
𝑖−1 = 𝑖 ∈ (𝐸2 , 𝐸1 ).
9.4 Lemma If ‖⋅‖1 and ‖⋅‖2 are two equivalent norms, then 𝐸1 ∶= (𝐸, ‖⋅‖1 ) is com-
plete if and only if 𝐸2 ∶= (𝐸, ‖⋅‖2 ) is complete.
Proof. Let (𝑥𝑛 ) be a sequence in 𝐸. Since ‖⋅‖1 and ‖⋅‖2 are equivalent there exists
𝑐, 𝐶 > 0 such that
𝑐‖𝑥𝑛 − 𝑥𝑚 ‖1 ≤ ‖𝑥𝑛 − 𝑥𝑚 ‖2 ≤ 𝐶‖𝑥𝑛 − 𝑥𝑚 ‖1
for all 𝑛, 𝑚 ∈ ℕ. Hence (𝑥𝑛 ) is a Cauchy sequence in 𝐸1 if and only if it is a Cauchy
sequence in 𝐸2 . Denote by 𝑖(𝑥) ∶= 𝑥 the identity map. If 𝑥𝑛 → 𝑥 in 𝐸1 , then 𝑥𝑛 → 𝑥
in 𝐸2 since 𝑖 ∈ (𝐸1 , 𝐸2 ) by Remark 9.3(d). Similarly 𝑥𝑛 → 𝑥 in 𝐸1 if 𝑥𝑛 → 𝑥 in 𝐸2
since 𝑖 ∈ (𝐸2 , 𝐸1 ).
Every linear operator between normed spaces induces a norm on its domain. We show
under what circumstances it is equivalent to the original norm.
9.5 Definition (Graph norm) Suppose that 𝐸, 𝐹 are normed spaces and 𝑇 ∈ Hom(𝐸, 𝐹 ).
We call
‖𝑢‖𝑇 ∶= ‖𝑢‖𝐸 + ‖𝑇 𝑢‖𝐹
the graph norm on 𝐸 associated with 𝑇 .
Let us now explain the term “graph norm.”
9.6 Remark From the linearity of 𝑇 it is rather evident that ‖⋅‖𝑇 is a norm on 𝐸. It is
called the “graph norm” because it really is a norm on the graph of 𝑇 . The graph of 𝑇
is the set
graph(𝑇 ) ∶= {(𝑢, 𝑇 𝑢) ∶ 𝑢 ∈ 𝐸} ⊂ 𝐸 × 𝐹 .
By the linearity of 𝑇 that graph is a linear subspace of 𝐸 × 𝐹 , and ‖⋅‖𝑇 is a norm,
making graph(𝑇 ) into a normed space. Hence the name graph norm.
Since ‖𝑢‖𝐸 ≤ ‖𝑢‖𝐸 +‖𝑇 𝑢‖𝐹 = ‖𝑢‖𝑇 for all 𝑢 ∈ 𝐸 the graph norm of any linear operator
is stronger than the norm on 𝐸. We have equivalence if and only if 𝑇 is bounded!
9.7 Proposition Let 𝑇 ∈ Hom(𝐸, 𝐹 ) and denote by ‖⋅‖𝑇 the corresponding graph
norm on 𝐸. Then 𝑇 ∈ (𝐸, 𝐹 ) if and only if ‖⋅‖𝑇 is equivalent to ‖⋅‖𝐸 .
Proof. If 𝑇 ∈ (𝐸, 𝐹 ), then
( )
‖𝑢‖𝐸 ≤ ‖𝑢‖𝑇 = ‖𝑢‖𝐸 + ‖𝑇 𝑢‖𝐹 ≤ ‖𝑢‖𝐸 + ‖𝑇 ‖(𝐸,𝐹 ) ‖𝑢‖𝐸 = 1 + ‖𝑇 ‖(𝐸,𝐹 ) ‖𝑢‖𝐸
for all 𝑢 ∈ 𝐸, showing that ‖⋅‖𝑇 and ‖⋅‖𝐸 are equivalent. Now assume that the two
norms are equivalent. Hence there exists 𝐶 > 0 such that ‖𝑢‖𝑇 ≤ 𝐶‖𝑢‖𝐸 for all 𝑢 ∈ 𝐸.
Therefore,
‖𝑇 𝑢‖𝐹 ≤ ‖𝑢‖𝐸 + ‖𝑇 𝑢‖𝐹 = ‖𝑢‖𝑇 ≤ 𝐶‖𝑢‖𝐸
for all 𝑢 ∈ 𝐸. Hence 𝑇 ∈ (𝐸, 𝐹 ) by Theorem 8.3.
32
10 Finite Dimensional Normed Spaces
In the previous section we did not make any assumption on the dimension of a vector
space. We prove that all norms on such spaces are equivalent.
10.1 Theorem Suppose that 𝐸 is a finite dimensional vector space. Then all norms on
𝐸 are equivalent. Moreover, 𝐸 is complete with respect to every norm.
Proof. Suppose that dim 𝐸 = 𝑁. Given a basis (𝑒1 , … , 𝑒𝑁 ), for every 𝑥 ∈ 𝐸 there
exist unique scalars 𝜉1 , … , 𝜉𝑁 ∈ 𝕂 such that
∑
𝑁
𝑥= 𝜉𝑖 𝑒𝑖 .
𝑖=1
∑
𝑁
𝑇 𝜉 ∶= 𝜉𝑖 𝑒𝑖 (10.1)
𝑖=1
(∑ )1∕2
‖∑ ∑
𝑁 𝑁 𝑁
‖
‖𝑇 𝜉‖𝐸 = ‖ 𝜉𝑖 𝑒𝑖 ‖ ≤ |𝜉𝑖 |‖𝑒𝑖 ‖𝐸 ≤ ‖𝑒𝑖 ‖2𝐸 |𝜉|2 = 𝐶|𝜉|2
‖ ‖𝐸
𝑛=1 𝑛=1 𝑛=1
(∑ )1∕2
𝑁
for all 𝜉 ∈ 𝕂 𝑁
if we set 𝐶 ∶= 𝑛=1
‖𝑒𝑖 ‖2𝐸 . Hence, 𝑇 ∈ (𝕂𝑁 , 𝐸). By the
continuity of a norm the map 𝜉 → ‖𝑇 𝜉‖𝐸 is a continuous map from 𝕂𝑁 to ℝ. In
particular it is continuous on the unit sphere 𝑆 = {𝜉 ∈ 𝕂𝑁 ∶ |𝜉|2 = 1}. Clearly 𝑆 is
a compact subset of 𝕂𝑁 . We know from Theorem 5.7 that continuous functions attain
a minimum on such a set. Hence there exists 𝛽 ∈ 𝑆 such that ‖𝑇 𝛽‖𝐸 ≤ ‖𝑇 𝜉‖𝐸 for
all 𝜉 ∈ 𝑆. Since ‖𝛽‖ = 1 ≠ 0 and 𝑇 is an isomorphism, property (i) of a norm (see
Definition 6.1) implies that 𝑐 ∶= ‖𝑇 𝛽‖𝐸 > 0. If 𝜉 ≠ 0, then since 𝜉∕|𝜉|2 ∈ 𝑆,
‖ 𝜉 ‖
𝑐 ≤ ‖𝑇 ‖ .
‖ |𝜉|2 ‖𝐸
Now by the linearity of 𝑇 and property (ii) of a norm 𝑐|𝜉|2 ≤ ‖𝑇 𝜉‖𝐸 and therefore
|𝑇 −1 𝑥|2 ≤ 𝑐 −1 ‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸. Hence 𝑇 −1 ∈ (𝐸, 𝕂𝑁 ) as claimed. We finally need
to prove completeness. Given a Cauchy sequence in 𝐸 with respect to ‖⋅‖𝐸 we have
from what we just proved that
|𝑇 −1 𝑥𝑛 − 𝑇 −1 𝑥𝑚 |2 ≤ 𝑐‖𝑥𝑛 − 𝑥𝑚 ‖𝐸 ,
33
showing that (𝑇 −1 𝑥𝑛 ) is a Cauchy sequence in 𝕂𝑁 . By the completeness of 𝕂𝑁 we
have 𝑇 −1 𝑥𝑛 → 𝜂 in 𝕂𝑁 . By continuity of 𝑇 proved above we get 𝑥𝑛 → 𝑇 𝜂, so (𝑥𝑛 )
converges. Since the above arguments work for every norm on 𝐸, this shows that 𝐸 is
complete with respect to every norm.
There are some useful consequences to the above theorem. The first is concerned with
finite dimensional subspaces of an arbitrary normed space.
10.2 Corollary Every finite dimensional subspace of a normed space 𝐸 is closed and
complete in 𝐸.
Proof. If 𝐹 is a finite dimensional subspace of 𝐸, then 𝐹 is a normed space with the
norm induced by the norm of 𝐸. By the above theorem 𝐹 is complete with respect to
that norm, so in particular it is closed in 𝐸.
The second shows that any linear operator on a finite dimensional normed space is
continuous.
We finally prove a counterpart to the Heine-Borel Theorem (Theorem 4.5) for general
finite dimensional normed spaces. In the next section we will show that the converse
is true as well, providing a topological characterisation of finite dimensional normed
spaces.
34
Proof. Fix an arbitrary 𝑥 ∈ 𝐸 ⧵ 𝑀 which exists since 𝑀 is a proper subspace of 𝐸.
As 𝑀 is closed dist(𝑥, 𝑀) ∶= 𝛼 > 0 as otherwise 𝑥 ∈ 𝑀 ̄ = 𝑀. Let 𝜀 ∈ (0, 1) be
arbitrary and note that (1 − 𝜀) > 1. Hence by definition of an infimum there exists
−1
𝑚𝜀 ∈ 𝑀 such that
𝛼
‖𝑥 − 𝑚𝜀 ‖ ≤ . (11.1)
1−𝜀
We define 𝑥 − 𝑚𝜀
𝑥𝜀 ∶= .
‖𝑥 − 𝑚𝜀 ‖
Then clearly ‖𝑥𝜀 ‖ = 1 and by (11.1) we have
‖ 𝑥 − 𝑚𝜀 ‖ 1 ‖ ( )‖
‖𝑥𝜀 − 𝑚‖ = ‖ − 𝑚‖ = ‖𝑥 − 𝑚𝜀 + ‖𝑥 − 𝑚𝜀 ‖𝑚 ‖
‖ ‖𝑥 − 𝑚𝜀 ‖ ‖ ‖𝑥 − 𝑚𝜀 ‖ ‖ ‖
1 − 𝜀‖ ( )‖
≥ ‖𝑥 − 𝑚𝜀 + ‖𝑥 − 𝑚𝜀 ‖𝑚 ‖
𝛼 ‖ ‖
for all 𝑚 ∈ 𝑀. As 𝑚𝜀 ∈ 𝑀 and 𝑀 is a subspace of 𝐸 we clearly have
𝑚𝜀 + ‖𝑥 − 𝑚𝜀 ‖𝑚 ∈ 𝑀
𝑀1 ⊊ 𝑀2 ⊊ 𝑀3 ⊊ ⋯ ⊊ 𝑀𝑛 ⊊ 𝑀𝑛+1
for all 𝑛 ∈ ℕ then there exist 𝑚𝑛 ∈ 𝑀𝑛 such that ‖𝑚𝑛 ‖ = 1 and dist(𝑚𝑛 , 𝑀𝑛−1 ) ≥ 1∕2
for all 𝑛 ∈ ℕ. Likewise, if
𝑀1 ⊋ 𝑀2 ⊋ 𝑀3 ⊋ ⋯ ⊋ 𝑀𝑛 ⊋ 𝑀𝑛+1
for all 𝑛 ∈ ℕ, then there exist 𝑚𝑛 ∈ 𝑀𝑛 such that ‖𝑚𝑛 ‖ = 1 and dist(𝑚𝑛 , 𝑀𝑛+1 ) ≥ 1∕2
for all 𝑛 ∈ ℕ.
Proof. Consider the first case. As 𝑀𝑛−1 is a proper closed subspace of 𝑀𝑛 we can
apply Theorem 11.1 and select 𝑚𝑛 ∈ 𝑀𝑛 such that dist(𝑚𝑛 , 𝑀𝑛−1 ) ≥ 1∕2. Doing so
inductively for all 𝑛 ∈ ℕ we get the required sequence (𝑚𝑛 ). In the second case we
proceed similarly: There exists 𝑚1 ∈ 𝑀1 such that dist(𝑚1 , 𝑀2 ) ≥ 1∕2. Next choose
𝑚2 ∈ 𝑀2 such that dist(𝑚2 , 𝑀3 ) ≥ 1∕2, and so on.
With the above we are able to give a topological characterisation of finite and infinite
dimensional spaces.
11.3 Theorem A normed space 𝐸 is finite dimensional if and only if the unit sphere
𝑆 = {𝑥 ∈ 𝐸 ∶ ‖𝑥‖ = 1} is compact.
35
Proof. First assume that dim 𝐸 = 𝑁 < ∞. Since the unit sphere is closed and
bounded, by the general Heine-Borel Theorem (Corollary 10.4) it is compact. Now
suppose that 𝐸 is infinite dimensional. Then there exists a countable linearly indepen-
dent set {𝑒𝑛 ∶ 𝑛 ∈ ℕ}. We set 𝑀𝑛 ∶= span{𝑒𝑘 ∶ 𝑘 = 1, … , 𝑛}. Clearly dim 𝑀𝑛 = 𝑛
and thus by Corollary 10.2 𝑀𝑛 is closed for all 𝑛 ∈ ℕ 𝑀𝑛 . As dim 𝑀𝑛+1 > dim 𝑀𝑛 the
sequence (𝑀𝑛 ) satisfies the assumptions of Corollary 11.2. Hence there exist 𝑚𝑛 ∈ 𝑀𝑛
such that ‖𝑚𝑛 ‖ = 1 and dist(𝑚𝑛 , 𝑀𝑛−1 ) ≥ 1∕2 for all 𝑛 ∈ ℕ. However, this implies that
‖𝑚𝑛 − 𝑚𝑘 ‖ ≥ 1∕2 whenever 𝑛 ≠ 𝑘, showing that there is a sequence in 𝑆 which does
not have a convergent subsequence. Hence 𝑆 cannot be compact, completing the proof
of the theorem.
11.4 Corollary Let 𝐸 be a normed vector space. Then the following assertions are
equivalent:
(i) dim 𝐸 < ∞;
12 Quotient Spaces
Consider a vector space 𝐸 and a subspace 𝐹 . We define an equivalence relation ∼
between elements 𝑥, 𝑦 in 𝐸 by 𝑥 ∼ 𝑦 if and only if 𝑥 − 𝑦 ∈ 𝐹 . Denote by [𝑥] the
equivalence class of 𝑥 ∈ 𝐸 and set
As you probably know from Algebra, this is called the quotient space of 𝐸 modulo 𝐹 .
That quotient space is a vector space over 𝕂 if we define the operations
[𝑥] + [𝑦] ∶= [𝑥 + 𝑦]
𝛼[𝑥] ∶= [𝛼𝑥]
36
for all 𝑥, 𝑦 ∈ 𝐸 and 𝛼 ∈ 𝕂. It is easily verified that these operations are well defined.
If 𝐸 is a normed space we would like to show that 𝐸∕𝐹 is a normed space with norm
This is a good definition since then ‖[𝑥]‖𝐸∕𝐹 ≤ ‖𝑥‖𝐸 for all 𝑥 ∈ 𝐸, that is, the natural
projection 𝐸 → 𝐸∕𝐹 , 𝑥 → [𝑥] is continuous. Geometrically, ‖[𝑥]‖𝐸∕𝐹 is the distance
of the affine subspace [𝑥] = 𝑥 + 𝐹 from the origin, or equivalently the distance between
the affine spaces 𝐹 and 𝑥 + 𝐹 as Figure 12.1 shows in the situation of two dimensions.
Unfortunately, ‖⋅‖𝐸∕𝐹 is not always a norm, but only if 𝐹 is a closed subspace of 𝐸.
[𝑥] = 𝑥 + 𝐹
𝑥
‖𝑥‖𝐸∕𝐹
𝐹
12.1 Proposition Let 𝐸 be a normed space. Then 𝐸∕𝐹 is a normed space with norm
(12.1) if and only if 𝐹 is a closed subspace of 𝐸.
Proof. Clearly ‖[𝑥]‖𝐸∕𝐹 ≥ 0 for all 𝑥 ∈ 𝐸 and ‖[𝑥]‖𝐸∕𝐹 = 0 if [𝑥] = [0], that is,
𝑥 ∈ 𝐹 . Now suppose that ‖[𝑥]‖𝐸∕𝐹 = 0. We want to show that then [𝑥] = [0] if and
only if 𝐹 is closed. First suppose that 𝐹 is closed. If ‖[𝑥]‖𝐸∕𝐹 = 0, then by definition
there exist 𝑧𝑛 ∈ 𝐹 with ‖𝑥 − 𝑧𝑛 ‖ → 0. Hence 𝑧𝑛 → 𝑥, and since 𝐹 is closed 𝑥 ∈ 𝐹 .
But then [𝑥] = [0] proving what we want. Suppose now 𝐹 is not closed. Then there
exists a sequence 𝑧𝑛 ∈ 𝐹 with 𝑧𝑛 → 𝑥 and 𝑥 ∉ 𝐹 . Hence [𝑥] ≠ [0], but
0 ≤ ‖[𝑥]‖𝐸∕𝐹 ≤ lim ‖𝑥 − 𝑧𝑛 ‖ = 0,
𝑛→∞
that is ‖[𝑥]‖𝐸∕𝐹 = 0 even though [𝑥] ≠ [0]. Hence (12.1) does not define a norm. The
other properties of a norm are valid no matter whether 𝐹 is closed or not. First note
that for 𝛼 ∈ 𝕂 and 𝑥 ∈ 𝐸
37
so ‖𝛼[𝑥]‖𝐸∕𝐹 ≥ |𝛼|‖𝑥‖𝐸∕𝐹 , showing that ‖𝛼[𝑥]‖𝐸∕𝐹 = |𝛼|‖𝑥‖𝐸∕𝐹 . Finally let 𝑥, 𝑦 ∈ 𝐸
and fix 𝜀 > 0 arbitrary. By definition of the quotient norm there exist 𝑧, 𝑤 ∈ 𝐹 such
that ‖𝑥 − 𝑧‖𝐸 ≤ ‖[𝑥]‖𝐸∕𝐹 + 𝜀 and ‖𝑦 − 𝑤‖𝐸 ≤ ‖[𝑦]‖𝐸∕𝐹 + 𝜀. Hence
12.3 Theorem Suppose 𝐸 is a Banach space and 𝐹 a closed subspace. Then 𝐸∕𝐹 is
a Banach space with respect to the quotient norm.
Proof. The only thing left to prove is that 𝐸∕𝐹 is complete with respect to the quo-
tient norm. We use the characterisation of completeness of a normed space given in
∑∞
Theorem 6.8. Hence let 𝑛=1 [𝑥𝑛 ] be an absolutely convergent series in 𝐸∕𝐹 , that is,
∑
∞
‖[𝑥𝑛 ]‖𝐸∕𝐹 ≤ 𝑀 < ∞.
𝑛=1
By definition of the quotient norm, for each 𝑛 ∈ ℕ there exists 𝑧𝑛 ∈ 𝐹 such that
1
‖𝑥𝑛 − 𝑧𝑛 ‖𝐸 ≤ ‖[𝑥𝑛 ]‖𝐸∕𝐹 + .
2𝑛
Hence,
∑
𝑚
∑
𝑚
∑𝑚
1
‖𝑥𝑛 − 𝑧𝑛 ‖𝐸 ≤ ‖[𝑥𝑛 ]‖𝐸∕𝐹 + 𝑛
≤𝑀 +2<∞
𝑛=1 𝑛=1 𝑛=1
2
∑∞
for all 𝑚 ∈ ℕ. This means that 𝑛=1 (𝑥𝑛 − 𝑧𝑛 ) is absolutely convergent and therefore
convergent by Theorem 6.8 and the assumption that 𝐸 be complete. We set
∑
∞
𝑠 ∶= (𝑥𝑛 − 𝑧𝑛 ).
𝑛=1
38
𝐸 ∶ 𝑇 𝑥 = 0} is a closed subspace of 𝐸. Hence 𝐸∕ ker 𝑇 is a normed space with
the quotient norm. We then define a linear operator 𝑇̂ ∶ 𝐸∕ ker 𝑇 → 𝐹 by setting
𝑇̂ [𝑥] ∶= 𝑇 𝑥
for all 𝑥 ∈ 𝐸. It is easily verified that this operator is well defined and linear. Moreover,
if we set 𝜋(𝑥) ∶= [𝑥], then by definition of the quotient norm ‖𝜋(𝑥)‖𝐸∕ ker 𝑇 ≤ ‖𝑥‖𝐸 ,
so 𝜋 ∈ (𝐸, 𝐸∕ ker 𝑇 ) with ‖𝜋‖(𝐸,𝐸∕ ker 𝑇 ) ≤ 1. Moreover, we have the factorisation
𝑇 = 𝑇̂ ◦𝜋,
𝑇
𝐸 𝐹
𝜋
𝑇̂
𝐸∕ ker 𝑇
12.4 Theorem Suppose that 𝐸, 𝐹 are normed spaces and that 𝑇 ∈ (𝐸, 𝐹 ). If 𝑇̂ , 𝜋
are defined as above, then 𝑇̂ ∈ (𝐸∕ ker 𝑇 , 𝐹 ), ‖𝑇 ‖(𝐸,𝐹 ) = ‖𝑇̂ ‖(𝐸∕ ker 𝑇 ,𝐹 ) and we
have the factorisation 𝑇 = 𝑇̂ ◦𝜋.
Proof. The only thing left to prove is that 𝑇̂ ∈ (𝐸∕ ker 𝑇 , 𝐹 ), and that ‖𝑇 ‖ ∶=
‖𝑇 ‖(𝐸,𝐹 ) = ‖𝑇̂ ‖(𝐸∕ ker 𝑇 ,𝐹 ) = ∶ ‖𝑇̂ ‖. First note that
Now by definition of the operator norm ‖𝑇̂ ‖ ≤ ‖𝑇 ‖ < ∞. In particular, 𝑇̂ ∈ (𝐸∕ ker 𝑇 , 𝐹 ).
To show equality of the operator norms observe that
by definition of the operator and quotient norms. Hence ‖𝑇 ‖ ≤ ‖𝑇̂ ‖, showing that
‖𝑇 ‖ = ‖𝑇̂ ‖.
39
40
Chapter III
Banach algebras and the Stone-Weierstrass
Theorem
13 Banach algebras
Some Banach spaces have an additional structure. For instance if we consider the space
of continuous functions 𝐶(𝐾) with 𝐾 compact, then we can multiply the functions as
well and define
(𝑓 𝑔)(𝑥) ∶= 𝑓 (𝑥)𝑔(𝑥)
for all 𝑥 ∈ 𝐾. The natural norm on 𝐶(𝐾) is the supremum norm, and we easily see
that
‖𝑓 𝑔‖∞ ≤ ‖𝑓 ‖∞ ‖𝑔‖∞ ,
which in particular implies that multiplication is continuous. Also, there is a neutral
element for multiplication, namely the constant function with value one. Vector spaces
with this additional multiplicative structure are called algebras. Here we have an addi-
tional topological structure. We introduce the following definition.
41
Young’s inequality implies that ‖𝑓 ∗ 𝑔‖1 ≤ ‖𝑓 ‖1 ‖𝑔‖1 , but there is no unity (see
MATH3969). Unity would correspond to the “delta function” which is really a measure,
not a function.
which is a power series in 𝑡 = (1 − 𝑥2 ), converging for |𝑡| < 1, that is, 𝑥 ∈ (−1, 1).
If we can prove that it converges absolutely and uniformly with respect to 𝑡 ∈ [−1, 1],
then the Weierstrass 𝑀-test shows that (14.1) converges absolutely and uniformly with
respect to 𝑥 ∈ [−1, 1]. Hence the sequence of partial sums of (14.1) provides a uniform
polynomial approximation of |𝑥| on [−1, 1].
converges absolutely and uniformly with respect to 𝑡 in the closed ball 𝐵(0, 1) ⊆ ℂ.
Proof. We know that (14.2) holds for 𝑡 in the open ball 𝐵(0, 1). As 𝛼 ∈ (0, 1) we have
( ) |( )|
𝛼 𝛼(𝛼 − 1) … (𝛼 − 𝑘 + 1) 𝑘−1 | 𝛼 |
= = (−1) | |
𝑘 𝑘! | 𝑘 |
| |
for all 𝑘 ≥ 1. Hence, for 𝑡 ∈ (0, 1) the series (14.2) can be rewritten in the form
∞ |( )|
∑ | 𝛼 | 𝑘
𝛼
(1 + 𝑡) = 1 − | |𝑡 . (14.3)
| 𝑘 |
𝑘=0 | |
42
Rearranging we obtain
𝑛 |( )| ∞ |( )|
∑ | 𝛼 | 𝑘 ∑| 𝛼 | 𝑘
| |𝑡 ≤ | | 𝑡 = 1 − (1 − 𝑡)𝛼 ≤ 1
| 𝑘 | | 𝑘 |
𝑘=0 | | 𝑘=0 | |
for all 𝑛 ∈ ℕ and all 𝑡 ∈ (0, 1). Letting 𝑡 ↑ 1 we see that
𝑛 |( )|
∑ | 𝛼 |
| |≤1
| 𝑘 |
𝑘=0 | |
for all 𝑛 ∈ ℕ. These are the partial sums of a series with non-negative terms and hence,
∞ |( )|
∑ | 𝛼 |
| |≤1
| 𝑘 |
𝑘=0 | |
converges. The Weierstrass 𝑀-Test now implies that (14.2) converges absolutely and
uniformly on 𝐵(0, 1).
We now state and prove the main result of this section about the density of sub-
algebras of 𝐶(𝐾). A sub-algebra is simply a subspace of 𝐶(𝐾) that is also a normed
algebra.
(ii) 𝐵 separates points, that is, for every pair of distinct points 𝑥, 𝑦 ∈ 𝐾 there exist
𝑓 ∈ 𝐵 such that 𝑓 (𝑥) ≠ 𝑓 (𝑦).
43
for all 𝑛 ∈ ℕ and all 𝑡 ∈ [−𝑛, 𝑛]. Let 𝑓 ∈ 𝐵. Since 𝐵 is an algebra 𝑝𝑛 ◦𝑓 ∈ 𝐵 for all
𝑛 ∈ ℕ, and that
| |
|(𝑝𝑛 ◦𝑓 )(𝑥) − |𝑓 (𝑥)|| < 1∕𝑛 (14.5)
| |
for all 𝑛 ≥ ‖𝑓 ‖∞ < ∞ and all 𝑥 ∈ 𝐾. Hence 𝑝𝑛 ◦𝑓 → |𝑓 | in 𝐶(𝐾), which means that
|𝑓 | ∈ 𝐵 for all 𝑓 ∈ 𝐵. As a consequence also
1( ) 1( )
max{𝑓 , 𝑔} = 𝑓 + 𝑔 + |𝑓 − 𝑔| and min{𝑓 , 𝑔} = 𝑓 + 𝑔 − |𝑓 − 𝑔| (14.6)
2 2
are in the closure of 𝐵 for all 𝑓 , 𝑔 ∈ 𝐵.
Fix now 𝑓 ∈ 𝐶(𝐾) and 𝑥, 𝑦 ∈ 𝐾 with 𝑥 ≠ 𝑦. By assumption (ii) there exists 𝑔 ∈ 𝐵
such that 𝑔(𝑥) ≠ 𝑔(𝑦). We set
( ) 𝑔(𝑧) − 𝑔(𝑥)
ℎ𝑥𝑦 (𝑧) ∶= 𝑓 (𝑥)1 + 𝑓 (𝑦) − 𝑓 (𝑥)
𝑔(𝑦) − 𝑔(𝑥)
for all 𝑧 ∈ 𝐾. As 1 ∈ 𝐵 by assumption we have ℎ𝑥𝑦 ∈ 𝐵 and ℎ𝑥𝑦 (𝑥) = 𝑓 (𝑥) and
ℎ𝑥𝑦 (𝑦) = 𝑓 (𝑦). Let now 𝜀 > 0 be fixed. Using that ℎ𝑥𝑦 and 𝑓 are continuous the sets
{ }
𝑈𝑥𝑦 ∶= 𝑧 ∈ 𝐾 ∶ ℎ𝑥𝑦 (𝑧) < 𝑓 (𝑧) + 𝜀
{ }
𝑉𝑥𝑦 ∶= 𝑧 ∈ 𝐾 ∶ ℎ𝑥𝑦 (𝑧) > 𝑓 (𝑧) − 𝜀
are open in 𝐾; see Theorem 5.3. Both sets are non-empty since ℎ𝑥𝑦 (𝑥) = 𝑓 (𝑥) and
ℎ𝑥𝑦 (𝑦) = 𝑓 (𝑦) and thus 𝑥, 𝑦 ∈ 𝑈𝑥𝑦 ∩ 𝑉𝑥𝑦 . The family (𝑈𝑥𝑦 )𝑥,𝑦∈𝐾 is clearly an open
cover of 𝐾. Fix now 𝑦 ∈ 𝐾. By the compactness of 𝐾 there exist finitely many points
⋃𝑚
𝑥𝑗 ∈ 𝐾, 𝑘 = 1, … , 𝑚, so that 𝐾 = 𝑗=1 𝑈𝑥𝑗 𝑦 . Applying (14.6) repeatedly we see that
𝑓𝑦 ∶= min ℎ𝑥𝑗 𝑦 ∈ 𝐵.
𝑗=1,…,𝑚
As 𝑉𝑦 is the intersection of finitely many open sets all containing 𝑦 it is open and non-
⋃𝓁
empty. Again by compactness of 𝐾 we can choose 𝑦1 , … , 𝑦𝓁 such that 𝐾 = 𝑗=1 𝑈𝑦𝑗 .
Again applying (14.6) we see that
ℎ ∶= max 𝑓𝑦 ∈ 𝐵,
𝑗=1,…,𝑚
and that
𝑓 (𝑧) − 𝜀 < ℎ(𝑧) < 𝑓 (𝑧) + 𝜀
for all 𝑧 ∈ 𝐾. In particular ‖𝑓 − ℎ‖∞ ≤ 𝜀. As ℎ ∈ 𝐵 and the argument above works
for any 𝜀 >, we conclude that 𝑓 ∈ 𝐵 as well.
44
(b) Let us now consider the case of 𝕂 = ℂ. Given 𝑓 ∈ 𝐶(𝐾, ℂ) = 𝐶(𝐾, ℝ) +
𝑖𝐶(𝐾, ℝ), assumption (iii) implies that real and imaginary parts given by
1 1
Re 𝑓 = (𝑓 + 𝑓̄) and Re 𝑓 = (𝑓 − 𝑓̄)
2 2𝑖
are both in 𝐵. Set 𝐵ℝ ∶= 𝐵 ∩ 𝐶(𝐾, ℝ). We can apply the real version proved already
to conclude that 𝐵ℝ = 𝐶(𝐾, ℝ). Hence, we also have 𝐵 = 𝐶(𝐾, ℂ).
14.4 Remark The Weierstrass approximation theorem does not apply to 𝐶(𝐾) if 𝐾
is a compact subset of ℂ containing some interior. In fact, polynomials are naturally
analytic funtions and a theorem from complex analysis asserts that the uniform limit
of analytic functions are analytic. Hence it is impossible to approximate an arbitrary
continuous function uniformly by polynomials unless it is analytic. All assumptions of
Theorem 14.2 are fulfilled except for the last, the one on the complex conjugates: if
𝑝(𝑧) is a polynomial, then its complex conjugate 𝑝(𝑧) is no longer a polynomial as 𝑧̄
cannot be writen as a polynomial in 𝑧.
45
46
Chapter IV
Hilbert Spaces
Hilbert spaces are in some sense a direct generalisation of finite dimensional Euclidean
spaces, where the norm has some geometric meaning and angles can be defined by
means of the dot product. The dot product can be used to define the norm and prove
many of its properties. Hilbert space theory is doing this in a similar fashion, where an
inner product is a map with properties similar to the dot product in Euclidean space. We
will emphasise the analogies and see how useful they are to find proofs in the general
context of inner product spaces.
(𝑢 ∣ 𝜆𝑣) = 𝜆(𝑢 ∣ 𝑣)
47
Next we give some examples of Banach and Hilbert spaces.
15.3 Examples (a) The space ℂ𝑁 equipped with the Euclidean scalar product given by
∑
𝑁
(𝑥 ∣ 𝑦) ∶= 𝑥 ⋅ 𝑦 = 𝑥𝑖 𝑦𝑖
𝑖=1
(𝑥 ∣ 𝑦)𝐴 ∶= 𝑥𝑇 𝐴𝑦̄
for all (𝑥𝑖 ), (𝑦𝑖 ) ∈ 𝓁2 . The series converges by Hölder’s inequality (Proposition 7.4).
(c) For 𝑢, 𝑣 ∈ 𝐿2 (𝑋) we let
(𝑢 ∣ 𝑣) ∶= 𝑢(𝑥)𝑣(𝑥) 𝑑𝑥.
∫𝑋
By Hölder’s inequality Proposition 7.7 the integral is finite and easily shown to be an
inner product.
The Euclidean norm on ℂ𝑁 is defined by means of the dot product, namely by ‖𝑥‖ =
√
𝑥 ⋅ 𝑥 for 𝑥 ∈ ℂ𝑁 . We make a similar definition in the context of general inner product
spaces.
15.4 Definition (induced norm) If 𝐸 is an inner product space with inner product (⋅ ∣
⋅) we define √
‖𝑢‖ ∶= (𝑢 ∣ 𝑢) (15.1)
for all 𝑢 ∈ 𝐸.
Note that from Remark 15.2 we always have (𝑥 ∣ 𝑥) ≥ 0, so ‖𝑥‖ is well defined. We
call ‖⋅‖ a “norm,” but at the moment we do not know whether it really is a norm in the
proper sense of Definition 6.1. We now want to work towards a proof that ‖⋅‖ is a norm
on 𝐸. On the way we look at some geometric properties of inner products and establish
the Cauchy-Schwarz inequality.
By the algebraic properties of the inner products in a space over ℝ and the definition
of the norm we get
‖𝑣 − 𝑢‖2 = ‖𝑢‖2 + ‖𝑣‖2 − 2(𝑢 ∣ 𝑣).
On the other hand, by the law of cosines we know that for vectors 𝑢, 𝑣 ∈ ℝ2
48
‖𝑣‖
𝜃 ‖𝑣 − 𝑢‖
‖𝑢‖
𝑢 ⋅ 𝑣 = ‖𝑢‖‖𝑣‖ cos 𝜃
and thus
|𝑢 ⋅ 𝑣| ≤ ‖𝑢‖‖𝑣‖.
The latter inequality has a counterpart in general inner product spaces. We give a proof
inspired by (but not relying on) the geometry in the plane. All arguments used purely
depend on the algebraic properties of an inner product and the definition of the induced
norm.
𝑛=𝑣−𝑝
𝑣 𝑢
(𝑢 ∣ 𝑣)
𝑝= 𝑢
‖𝑢‖2
49
of the norm we get
(𝑢 ∣ 𝑣)(𝑣 ∣ 𝑢) (𝑢 ∣ 𝑣)(𝑢 ∣ 𝑣)
0 ≤ ‖𝑛‖2 = 𝑣 ⋅ 𝑣 − 2 + (𝑢 ∣ 𝑢)
‖𝑢‖2 ‖𝑢‖4
|(𝑢 ∣ 𝑣)|2 |(𝑢 ∣ 𝑣)|2 |(𝑢 ∣ 𝑣)|2
= ‖𝑣‖2 − 2 + ‖𝑢‖ 2
= ‖𝑣‖2
− .
‖𝑢‖2 ‖𝑢‖4 ‖𝑢‖2
Therefore |(𝑢 ∣ 𝑣)|2 ≤ ‖𝑢‖2 ‖𝑣‖2 , and by taking square roots we find (15.2). Clearly
equality holds if and only if ‖𝑛‖ = 0, that is, if
(𝑢 ∣ 𝑣)
𝑣= 𝑢.
‖𝑢‖2
Hence we have equality in (15.2) if and only if 𝑢 and 𝑣 are linearly dependent. This
completes the proof of the theorem.
15.6 Corollary If 𝐸 is an inner product space and ‖⋅‖ the induced norm, then
‖𝑢‖ = sup |(𝑢 ∣ 𝑣)| = sup |(𝑢 ∣ 𝑣)|
‖𝑣‖≤1 ‖𝑣‖=1
for all 𝑢 ∈ 𝐸.
Proof. If 𝑢 = 0 the assertion is obvious, so assume that 𝑢 ≠ 0. If ‖𝑣‖ ≤ 1, then
|(𝑢 ∣ 𝑣)| ≤ ‖𝑢‖‖𝑣‖ = ‖𝑢‖ by the Cauchy-Schwarz inequality. Hence
‖𝑢‖ ≤ sup |(𝑢 ∣ 𝑣)|.
‖𝑣‖≤1
Choosing 𝑣 ∶= 𝑢∕‖𝑢‖ we have |(𝑢 ∣ 𝑣)| = ‖𝑢‖2 ∕‖𝑢‖ ≤ ‖𝑢‖, so equality holds in
the above inequality. Since the supremum over ‖𝑣‖ = 1 is larger or equal to that over
‖𝑣‖ ≤ 1, the assertion of the corollary follows.
Using the Cauchy-Schwarz inequality we can now prove that ‖⋅‖ is in fact a norm.
50
15.8 Convention Since every inner product induces a norm we will always assume that
an inner product space is a normed space with the norm induced by the inner product.
Once we have a norm we can talk about convergence and completeness. Note that not
every inner product space is complete, but those which are play a special role.
15.9 Definition (Hilbert space) An inner product space which is complete with re-
spect to the induced norm is called a Hilbert space.
The inner product is a map on 𝐸 × 𝐸. We show that this map is continuous with respect
to the induced norm.
Proof. If 𝑥𝑛 → 𝑥 and 𝑦𝑛 → 𝑦 in 𝐸 (with respect to the induced norm), then using the
Cauchy-Schwarz inequality
as 𝑛 → ∞. Note that we also use the continuity of the norm in the above argument to
conclude that ‖𝑦𝑛 ‖ → ‖𝑦‖ (see Theorem 6.4). Hence the inner product is continuous.
The lengths of the diagonals and edges of a parallelogram in the plane satisfy a rela-
tionship. The norm in an inner product space satisfies a similar relationship, called the
parallelogram identity. The identity will play an essential role in the next section.
15.11 Proposition (Parallelogram identity) Let 𝐸 be an inner product space and ‖⋅‖
the induced norm. Then
for all 𝑢, 𝑣 ∈ 𝐸.
Proof. By definition of the induced norm and the properties of an inner product
It turns out that the converse is true as well. More precisely, if a norm satisfies (15.3)
for all 𝑢, 𝑣 ∈ 𝐸, then there is an inner product inducing that norm (see [13, Section I.5]
for a proof).
51
16 Projections and Orthogonal Complements
In this section we discuss the existence and properties of “nearest point projections”
from a point onto a set, that is, the points that minimise the distance from a closed set
to a given point.
The meaning of 𝑃𝑀 (𝑥) is illustrated in Figure 16.1 for the Euclidean norm in the plane.
If the set is not convex, 𝑃𝑀 (𝑥) can consist of several points, if it is convex, it is precisely
one.
We now look at some example. First we look at subsets of ℝ𝑁 , and show that then
𝑃𝑀 (𝑥) is never empty.
16.3 Example Let 𝐸 ∶= 𝐶([0, 1]) with norm ‖𝑢‖ ∶= ‖𝑢‖∞ + ‖𝑢‖1 . We claim that 𝐸
is complete. Clearly ‖𝑢‖∞ ≤ ‖𝑢‖ for all 𝑢 ∈ 𝐸. Also
1 1
‖𝑢‖1 = |𝑢(𝑥)| 𝑑𝑥 ≤ ‖𝑢‖∞ 1 𝑑𝑥 = ‖𝑢‖∞
∫0 ∫0
52
for all 𝑥 ∈ 𝐸. Hence ‖⋅‖ is equivalent to the supremum norm ‖⋅‖∞ , so convergence
with respect to ‖⋅‖ is uniform convergence. By Section 7.4 and Lemma 9.4 the space
𝐸 is complete with respect to the norm ‖⋅‖. Now look at the subspace
In the light of the above example it is a non-trivial fact that 𝑃𝑀 (𝑥) is not empty in a
Hilbert space, at least if 𝑀 is closed and convex. By a convex set, as usual, we mean a
subset 𝑀 such that 𝑡𝑥 + (1 − 𝑡)𝑦 ∈ 𝑀 for all 𝑥, 𝑦 ∈ 𝑀 and 𝑡 ∈ [0, 1]. In other words,
if 𝑥, 𝑦 are in 𝑀, so is the line segment connecting them. The essential ingredient in the
proof is the parallelogram identity from Proposition 15.11.
Since 𝑀 is closed and 𝑥 ∉ 𝑀 we have 𝛼 > 0. From the parallelogram identity Propo-
sition 15.11 we get
53
If 𝑚1 , 𝑚2 ∈ 𝑀, then ‖𝑚𝑖 − 𝑥‖ ≥ 𝛼 for 𝑖 = 1, 2 and by the convexity of 𝑀 we have
(𝑚1 + 𝑚2 )∕2 ∈ 𝑀. Hence
‖ 𝑚 + 𝑚2 ‖
‖(𝑚1 − 𝑥) + (𝑚2 − 𝑥)‖ = ‖𝑚1 + 𝑚2 − 2𝑥‖ = 2‖ 1 − 𝑥‖ ≥ 2𝛼.
‖ 2 ‖
and by using the above
‖𝑚1 − 𝑚2 ‖2 ≤ 4𝛼 2 − 4𝛼 2 = 0.
This obviously implies that (𝑥𝑛 ) a bounded sequence in 𝐻, but since 𝐻 is not necessar-
ily finite dimensional, we cannot conclude it is converging without further investigation.
We show that (𝑥𝑛 ) is a Cauchy sequence and therefore converges by the completeness
of 𝐻. Fix now 𝜀 > 0. Since 𝛼 ≤ ‖𝑥𝑛 − 𝑥‖ → 𝛼 there exists 𝑛0 ∈ ℕ such that
𝛼 ≤ ‖𝑥𝑛 − 𝑥‖ ≤ 𝛼 + 𝜀
(i) 𝑚𝑥 = 𝑃𝑀 (𝑥);
54
𝑥
𝑥 − 𝑚𝑥
≥ 𝜋∕2
𝑚𝑥 = 𝑃𝑀 (𝑥)
𝑚 𝑚 − 𝑚𝑥
so 0 = 𝑃𝑀 (𝑥) as claimed.
16.6 Corollary Let 𝑀 be a closed subspace of the Hilbert space 𝐻. Then 𝑚𝑥 = 𝑃𝑀 (𝑥)
if and only if 𝑚𝑥 ∈ 𝑀 and (𝑥 − 𝑚𝑥 ∣ 𝑚) = 0 for all 𝑚 ∈ 𝑀. Moreover, 𝑃𝑀 ∶ 𝐻 → 𝑀
is linear.
Proof. By the above theorem 𝑚𝑥 = 𝑃𝑀 (𝑥) if and only if Re(𝑚𝑥 − 𝑥 ∣ 𝑚 − 𝑚𝑥 ) ≤ 0
for all 𝑚 ∈ 𝑀. Since 𝑀 is a subspace 𝑚 + 𝑚𝑥 ∈ 𝑀 for all 𝑚 ∈ 𝑀, so using 𝑚 + 𝑚𝑥
instead of 𝑚 we get that
Re(𝑚𝑥 − 𝑥 ∣ (𝑚 + 𝑚𝑥 ) − 𝑚𝑥 ) = Re(𝑚𝑥 − 𝑥 ∣ 𝑚) ≤ 0
55
𝑥
𝑥 − 𝑚𝑥
0
𝑚
𝑚𝑥 = 𝑃𝑀 (𝑥)
for all 𝑚 ∈ 𝑀. Hence again by what we proved 𝑃𝑀 (𝜆𝑥 + 𝜇𝑦) = 𝜆𝑃𝑀 (𝑥) + 𝜇𝑃𝑀 (𝑦),
showing that 𝑃𝑀 is linear.
We next connect the projections discussed above with the notion of orthogonal com-
plements.
16.8 Lemma Suppose 𝑀 is a non-empty subset of the inner product space 𝐻. Then
⟂
𝑀 ⟂ is a closed subspace of 𝐻 and 𝑀 ⟂ = 𝑀 = (span 𝑀)⟂ = (span 𝑀)⟂ .
Proof. If 𝑥, 𝑦 ∈ 𝑀 ⟂ and 𝜆, 𝜇 ∈ 𝕂, then
56
for all 𝑚 ∈ 𝑀. Hence 𝑥 ∈ 𝑀 ⟂ , showing that 𝑀 ⟂ is closed. We next show that 𝑀 ⟂ =
⟂ ⟂
𝑀 . Since 𝑀 ⊂ 𝑀 we have 𝑀 ⊂ 𝑀 ⟂ by definition the orthogonal complement.
Fix 𝑥 ∈ 𝑀 ⟂ and 𝑚 ∈ 𝑀. Then there exist 𝑚𝑛 ∈ 𝑀 with 𝑚𝑛 → 𝑚. By the continuity
of the inner product
(𝑥 ∣ 𝑚) = lim (𝑥 ∣ 𝑚𝑛 ) = lim 0 = 0.
𝑛→∞ 𝑛→∞
⟂ ⟂ ⟂
Hence 𝑥 ∈ 𝑀 and thus 𝑀 ⊃ 𝑀 ⟂ , showing that 𝑀 = 𝑀 ⟂ . Next we show that
𝑀 ⟂ = (span 𝑀)⟂ . Clearly (span 𝑀)⟂ ⊂ 𝑀 ⟂ since 𝑀 ⊂ span 𝑀. Suppose now that
𝑥 ∈ 𝑀 ⟂ and 𝑚 ∈ span 𝑀. Then there exist 𝑚𝑖 ∈ 𝑀 and 𝜆𝑖 ∈ 𝕂, 𝑖 = 1, … , 𝑛, such
∑𝑛
that 𝑚 = 𝑖=1 𝜆𝑖 𝑚𝑖 . Hence
∑
𝑛
(𝑥 ∣ 𝑚) = 𝜆𝑖 (𝑥 ∣ 𝑚𝑖 ) = 0,
𝑖=1
and thus 𝑥 ∈ (span 𝑀)⟂ . Therefore (span 𝑀)⟂ ⊃ 𝑀 ⟂ and so (span 𝑀)⟂ = 𝑀 ⟂
as claimed. The last assertion of the lemma follows by what we have proved above.
⟂ ⟂
Indeed we know that 𝑀 ⟂ = 𝑀 and that 𝑀 = (span 𝑀)⟂ .
We are now ready to prove the main result on orthogonal projections. It is one of the
most important and useful facts on Hilbert spaces.
(i) 𝐻 = 𝑀 ⊕ 𝑀 ⟂ ;
𝑥 = 𝑃𝑀 (𝑥) + (𝐼 − 𝑃𝑀 )(𝑥) ∈ 𝑀 + 𝑀 ⟂ ,
57
16.10 Remark The above theorem in particular implies that for every closed subspace
𝑀 of a Hilbert space 𝐻 there exists a closed subspace 𝑁 such that 𝐻 = 𝑀 ⊕ 𝑁.
We call 𝑀 a complemented subspace. The proof used the existence of a projection.
We know from Example 16.3 that projections onto closed subspaces do not necessarily
exist in a Banach space, so one may not expect every subspace of a Banach space to
be complemented. A rather recent result [9] shows that if every closed subspace of a
Banach space is complemented, then its norm is equivalent to a norm induced by an
inner product! Hence the above theorem provides a unique property of Hilbert spaces.
The above theorem can be used to prove some properties of orthogonal complements.
The first is a very convenient criterion for a subspace of a Hilbert space to be dense.
𝐻 = 𝑀 ⊕ 𝑀⟂
We finally use Theorem 16.9 to get a characterisation of the second orthogonal com-
plement of a set.
𝑀 ⟂⟂ ∶= (𝑀 ⟂ )⟂ = span 𝑀.
Proof. By Lemma 16.8 we have 𝑀 ⟂ = (span 𝑀)⟂ = (span 𝑀)⟂ . Hence by replacing
𝑀 by span 𝑀 we can assume without loss of generality that 𝑀 is a closed subspace of
𝐻. We have to show that 𝑀 = 𝑀 ⟂⟂ . Since (𝑥 ∣ 𝑚) = 0 for all 𝑥 ∈ 𝑀 and 𝑚 ∈ 𝑀 ⟂ we
have 𝑀 ⊂ 𝑀 ⟂⟂ . Set now 𝑁 ∶= 𝑀 ⟂ ∩ 𝑀 ⟂⟂ . Since 𝑀 is a closed subspace it follows
from Theorem 16.9 that 𝑀 ⟂⟂ = 𝑀 ⊕ 𝑁. By definition 𝑁 ⊂ 𝑀 ⟂ ∩ 𝑀 ⟂⟂ = {0}, so
𝑁 = {0}, showing that 𝑀 = 𝑀 ⟂⟂ .
17 Orthogonal Systems
In ℝ𝑁 , the standard basis or any other basis of mutually orthogonal vectors of length
one play a special role. We look at generalisations of such bases. Recall that two 𝑢, 𝑣
of an inner product space are called orthogonal if (𝑢 ∣ 𝑣) = 0.
17.1 Definition (orthogonal systems) Let 𝐻 be an inner product space with inner
product (⋅ ∣ ⋅) and induced norm ‖⋅‖. Let 𝑀 ⊂ 𝐻 be a non-empty subset.
58
(ii) 𝑀 is called an orthonormal system if it is an orthogonal system and ‖𝑢‖ = 1 for
all 𝑢 ∈ 𝑀.
(iii) 𝑀 is called a complete orthonormal system or orthonormal basis of 𝐻 if it is an
orthogonal system and span 𝑀 = 𝐻.
Note that the notion of orthogonal system depends on the particular inner product, so
we always have to say with respect to which inner product it is orthogonal.
which easily follow from using the standard addition theorems for sin(𝑚 ± 𝑛)𝑥 and
cos(𝑚 ± 𝑛)𝑥
59
We next show that orthogonal systems are linearly independent if we remove the zero
element. Recall that by definition an infinite set is linearly independent if every finite
subset is linearly independent. We also prove a generalisation of Pythagoras’ theorem.
17.3 Lemma (Pythagoras theorem) Suppose that 𝐻 is an inner product space and
𝑀 an orthogonal system in 𝐻. Then the following assertions are true:
(i) 𝑀 ⧵ {0} is linearly independent.
(ii) If (𝑥𝑛 ) is a sequence in 𝑀 with 𝑥𝑛 ≠ 𝑥𝑚 for 𝑛 ≠ 𝑚 and 𝐻 is complete, then
∑∞ ∑∞
𝑥 converges if and only if 𝑘=0 ‖𝑥𝑘 ‖2 converges. In that case
𝑘=0 𝑘
‖∑ ‖2 ∑
∞ ∞
‖ 𝑥𝑘 ‖ = ‖𝑥𝑘 ‖2 . (17.1)
‖ ‖
𝑘=0 𝑘=0
Proof. (i) We have to show that every finite subset of 𝑀 ⧵ {0} is linearly independent.
Hence let 𝑥𝑘 ∈ 𝑀 ⧵ {0}, 𝑘 = 1, … , 𝑛 be a finite number of distinct elements. Assume
that 𝜆𝑘 ∈ 𝕂 are such that
∑𝑛
𝜆𝑘 𝑥𝑘 = 0.
𝑘=0
Since 𝑥𝑚 ≠ 0 it follows that 𝜆𝑚 = 0 for all 𝑚 ∈ {0, … , 𝑛}, showing that 𝑀 ⧵ {0} is
linearly independent.
(ii) Let (𝑥𝑛 ) be a sequence in 𝑀 with 𝑥𝑛 ≠ 𝑥𝑚 . (We only look at the case of
an infinite set because otherwise there are no issues on convergence). We set 𝑠𝑛 ∶=
∑𝑛 ∑𝑛
𝑥 and 𝑡𝑛 ∶= 𝑘=1 ‖𝑥𝑘 ‖2 the partial sums of the series under consideration. If
𝑘=1 𝑘
1 ≤ 𝑚 < 𝑛, then by the orthogonality
(∑ )
‖∑ | ∑
𝑛 𝑛 𝑛
‖2
‖𝑠𝑛 − 𝑠𝑚 ‖ = ‖
2
𝑥𝑘 ‖ = 𝑥𝑘 | 𝑥𝑗
‖ ‖ |
𝑘=𝑚+1 𝑘=𝑚+1 𝑗=𝑚+1
∑ ∑
𝑛 𝑛
∑
𝑛
= (𝑥𝑘 ∣ 𝑥𝑗 ) = ‖𝑥𝑘 ‖2 = |𝑡𝑛 − 𝑡𝑚 |.
𝑘=𝑚+1 𝑗=𝑚+1 𝑘=𝑚+1
60
If we do not sum over the full standard basis we may only get an inequality, namely
∑
𝑚
∑
𝑛
|(𝑥 ∣ 𝑒𝑖 )| ≤ 2
|(𝑥 ∣ 𝑒𝑖 )|2 ≤ ‖𝑥‖2 .
𝑘=1 𝑘=1
17.4 Theorem (Bessel’s inequality) Let 𝐻 be an inner product space and 𝑀 an or-
thonormal system in 𝐻. Then
∑
|(𝑥 ∣ 𝑚)|2 ≤ ‖𝑥‖2 (17.3)
𝑚∈𝑀
‖∑ ‖2 ∑ ∑
𝑛 𝑛 𝑛
‖ (𝑥 ∣ 𝑚𝑘 )𝑚𝑘 ‖ = |(𝑥 ∣ 𝑚𝑘 )|2 ‖𝑚𝑘 ‖2 = |(𝑥 ∣ 𝑚𝑘 )|2 .
‖ ‖
𝑘=1 𝑘=1 𝑘=1
We expect the norm of the projection to be smaller than the norm of ‖𝑥‖. To see that
we use the properties of the inner product and the above identity to get
‖ ∑
𝑛
‖2 ‖∑
𝑛
‖2
0 ≤ ‖𝑥 − (𝑥 ∣ 𝑚𝑘 )𝑚𝑘 ‖ = ‖𝑥‖ + ‖ (𝑥 ∣ 𝑚𝑘 )𝑚𝑘 ‖
2
‖ ‖ ‖ ‖
𝑘=1 𝑘=1
∑
𝑛
∑ 𝑛
− (𝑥 ∣ 𝑚𝑘 )(𝑥 ∣ 𝑚𝑘 ) − (𝑥 ∣ 𝑚𝑘 )(𝑚𝑘 ∣ 𝑥)
𝑘=1 𝑘=1
∑
𝑛
∑
𝑛
= ‖𝑥‖2 + |(𝑥 ∣ 𝑚𝑘 )|2 − 2 |(𝑥 ∣ 𝑚𝑘 )|2
𝑘=1 𝑘=1
∑ 𝑛
= ‖𝑥‖2 − |(𝑥 ∣ 𝑚𝑘 )|2 .
𝑘=1
61
Hence we have shown that ∑
|(𝑥 ∣ 𝑚)|2 ≤ ‖𝑥‖2
𝑚∈𝑁
for every finite set 𝑁 ⊂ 𝑀. Taking the supremum over all such finite sets (17.3)
follows. To prove the second assertion note that for every given 𝑥 ∈ 𝐻 the sets 𝑀𝑛 ∶=
{𝑚 ∈ 𝑀 ∶ |(𝑥 ∣ 𝑚)| ≥ 1∕𝑛} is finite for every 𝑛 ∈ ℕ as otherwise (17.3) could not be
true. Since countable unions of finite sets are countable, the set
⋃
{𝑚 ∈ 𝑀 ∶ (𝑥 ∣ 𝑚) ≠ 0} = 𝑀𝑛
𝑛∈ℕ
is countable as claimed.
∑
𝑛
(𝑥 ∣ 𝑒𝑘 )𝑒𝑘
𝑘=1
62
is at most countable. Hence 𝑀𝑥 is finite or its elements can be enumerated. If 𝑀𝑥 is
finite (18.1) makes perfectly good sense. Hence let us assume that 𝑚𝑘 , 𝑘 ∈ ℕ is an
enumeration of 𝑀𝑥 . Hence, rather than (18.1), we could write
∑
∞
(𝑥 ∣ 𝑚𝑘 )𝑚𝑘 .
𝑘=0
This does still not solve all our problems, because the limit of the series may depend
on the particular enumeration chosen. The good news is that this is not the case, and
that the series is unconditionally convergent, that is, the series converges and for every
bijection 𝜎 ∶ ℕ → ℕ we have
∑
∞
∑
∞
(𝑥 ∣ 𝑚𝑘 )𝑚𝑘 = (𝑥 ∣ 𝑚𝜎(𝑘) )𝑚𝜎(𝑘) .
𝑘=0 𝑘=0
Recall that the series on the right hand side is called a rearrangement of the series on
the left. We now show that (18.1) is actually a projection, not onto span 𝑀, but onto
its closure.
18.1 Theorem Suppose that 𝑀 is an orthonormal system in a Hilbert space 𝐻 and set
∑∞
𝑁 ∶= span 𝑀. Let 𝑥 ∈ 𝐻 and 𝑚𝑘 , 𝑘 ∈ ℕ an enumeration of 𝑀𝑥 . Then 𝑘=0 (𝑥 ∣ 𝑚𝑘 )𝑚𝑘
is unconditionally convergent, and
∑
∞
𝑃𝑁 (𝑥) = (𝑥 ∣ 𝑚𝑘 )𝑚𝑘 , (18.3)
𝑘=0
Since the series is convergent, the continuity of the inner product shows that
∑
∞
(𝑦 − 𝑥 ∣ 𝑚) = lim 𝑠𝑛 (𝑚) = (𝑥 ∣ 𝑚𝑘 )(𝑚𝑘 ∣ 𝑚) − (𝑥 ∣ 𝑚)
𝑛→∞
𝑘=0
63
exists for all 𝑚 ∈ 𝑀. If 𝑚 ∈ 𝑀𝑥 , that is, 𝑚 = 𝑚𝑗 for some 𝑗 ∈ ℕ, then by the
orthogonality
(𝑦 − 𝑥 ∣ 𝑚) = (𝑥 ∣ 𝑚𝑗 ) − (𝑥 ∣ 𝑚𝑗 ) = 0.
If 𝑚 ∈ 𝑀 ⧵ 𝑀𝑥 , then (𝑥 ∣ 𝑚) = (𝑚𝑘 ∣ 𝑚) = 0 for all 𝑘 ∈ ℕ by definition of 𝑀𝑥 and the
orthogonality. Hence again (𝑦 − 𝑥 ∣ 𝑚) = 0, showing that 𝑦 − 𝑥 ∈ 𝑀 ⟂ . By Lemma 16.8
⟂
it follows that 𝑦−𝑥 ∈ span 𝑀 . Now Corollary 16.6 implies that 𝑦 = 𝑃𝑁 (𝑥) as claimed.
Since we have worked with an arbitrary enumeration of 𝑀𝑥 and 𝑃𝑁 (𝑥) is independent
of that enumeration, it follows that the series is unconditionally convergent.
We have just shown that (18.3) is unconditionally convergent. For this reason we can
make the following definition, giving sense to (18.1).
18.2 Definition (Fourier series) Let 𝑀 be an orthonormal system in the Hilbert space
𝐻. If 𝑥 ∈ 𝐻 we call (𝑥 ∣ 𝑚), 𝑚 ∈ 𝑀, the Fourier coefficients of 𝑥 with respect to 𝑀.
Given an enumeration 𝑚𝑘 , 𝑘 ∈ ℕ of 𝑀𝑥 as defined in (18.2) we set
∑ ∑
∞
(𝑥 ∣ 𝑚)𝑚 ∶= (𝑥 ∣ 𝑚𝑘 )𝑚𝑘
𝑚∈𝑀 𝑘=0
and call it the Fourier series of 𝑥 with respect to 𝑀. (For convenience here we let
𝑚𝑘 = 0 for 𝑘 larger than the cardinality of 𝑀𝑥 if it is finite.)
With the above definition, Theorem 18.1 shows that
∑
(𝑥 ∣ 𝑚)𝑚 = 𝑃𝑁 (𝑥)
𝑚∈𝑀
64
if (ii) holds, so (iii) follows.
(iii)⇒(i): Let 𝑁 ∶= span 𝑀 and fix 𝑥 ∈ 𝑁 ⟂ . By assumption, Theorem 16.9
and 18.1 as well as Lemma 17.3 we have
‖∑ ‖2 ∑
0 = ‖𝑃𝑁 (𝑥)‖2 = ‖ (𝑥 ∣ 𝑚)𝑚‖ = |(𝑥 ∣ 𝑚)|2 = ‖𝑥‖2 .
‖ ‖
𝑚∈𝑀 𝑚∈𝑀
⟂
Hence 𝑥 = 0, showing that span 𝑀 = {0}. By Corollary 16.11 span 𝑀 = 𝐻, that is,
𝑀 is complete, proving (i).
We next provide the connection of the above “abstract Fourier series” to the “classical”
Fourier series you may have seen elsewhere. To do so we look at the expansions with
respect to the orthonormal systems considered in Example 17.2.
18.4 Example (a) Let 𝑒𝑖 be the standard basis in 𝕂𝑁 . The Fourier “series” of 𝑥 ∈ 𝕂𝑁
with respect to 𝑒𝑖 is
∑
𝑁
𝑥= (𝑥 ∣ 𝑒𝑖 )𝑒𝑖 .
𝑖=1
Of course we do not usually call this a “Fourier series” but say 𝑥𝑖 ∶= (𝑥 ∣ 𝑒𝑖 ) are the
components of the vector 𝑥 and the above sum the representation of 𝑥 with respect to
the basis 𝑒𝑖 . The example should just illustrate once more the parallels of Hilbert space
theory to various properties of Euclidean spaces.
(b) The Fourier coefficients of 𝑢 ∈ 𝐿2 ((−𝜋, 𝜋), ℂ) with respect to the orthonormal
system
1 𝑖𝑛𝑥
√ 𝑒 , 𝑛 ∈ ℤ,
2𝜋
are given by
𝜋
1
𝑐𝑛 ∶= √ 𝑒−𝑖𝑛𝑥 𝑢(𝑥) 𝑑𝑥.
∫
2𝜋 −𝜋
Hence the Fourier series of 𝑢 with respect to the above system is
∑ ∑ 1
𝜋
𝑢= 𝑐𝑛 𝑒𝑖𝑛𝑥 = √ 𝑒−𝑖𝑛𝑥 𝑢(𝑥) 𝑑𝑥𝑒𝑖𝑛𝑥 .
𝑛∈ℤ 𝑛∈ℤ
∫
2𝜋 −𝜋
This is precisely the complex form of the classical Fourier series of 𝑢. As the orthonor-
mal system under considerationis complete, our theory tells us that the series converges
in 𝐿2 ((−𝜋, 𝜋), ℂ), but we do not get any information on pointwise or uniform conver-
gence.
(c) We now look at 𝑢 ∈ 𝐿2 ((−𝜋, 𝜋), ℝ) and its expansion with respect to the or-
thonormal system given by
1 1 1
√ , √ cos 𝑛𝑥, √ sin 𝑛𝑥, 𝑛 ∈ ℕ ⧵ {0}.
2𝜋 𝜋 𝜋
65
The Fourier coefficients are
𝜋
1
𝑎0 = √ 𝑢(𝑥) 𝑑𝑥
2𝜋 ∫−𝜋
𝜋
1
𝑎𝑛 = √ 𝑢(𝑥) cos 𝑛𝑥 𝑑𝑥
𝜋 ∫−𝜋
𝜋
1
𝑏𝑛 = √ 𝑢(𝑥) sin 𝑛𝑥 𝑑𝑥
𝜋 ∫−𝜋
Hence the Fourier series with respect to the above system is
∑
∞
𝑢 = 𝑎0 + (𝑎𝑛 cos 𝑛𝑥 + 𝑏𝑛 sin 𝑛𝑥),
𝑛=0
18.5 Remark One can also get orthonormal systems from any finite or countable set of
linearly independent elements of an inner product space by means of the Gram-Schmidt
orghogonalisation process as seen in second year algebra.
We have mentioned the possibility of uncountable orthonormal systems or bases. They
can occur, but in practice all orthogonal bases arising from applications (like partial
differential equations) are countable. Recall that a metric space is separable if it has a
countable dense subset.
18.6 Theorem A Hilbert space is separable if and only if it has a countable orthonor-
mal basis.
Proof. If the space 𝐻 is finite dimensional and 𝑒𝑖 , 𝑖 = 1, … 𝑁, is an orthonormal basis
of 𝐻, then the set
{∑
𝑁 }
spanℚ {𝑒1 , … , 𝑒𝑁 } ∶= 𝜆𝑘 𝑒𝑘 ∶ 𝜆𝑘 ∈ ℚ(+𝑖ℚ)
𝑘=1
66
span{𝑒1 , … , 𝑒𝑁 }. Then dim 𝐻𝑁 = 𝑁 and by what we just proved, 𝐻𝑁 is separable.
Since countable unions of countable sets are countable it follows that countable unions
of separable sets are separable. Hence
⋃
span 𝑀 = 𝐻𝑁
𝑁∈ℕ
is separable. Since 𝑀 is complete span 𝑀 is dense. Hence any dense subset of span 𝑀
is dense in 𝐻 as well, proving that 𝐻 is separable. Assume now that 𝐻 is a sep-
arable Hilbert space and let 𝐷 ∶= {𝑥𝑘 ∶ 𝑘 ∈ ℕ} be a dense subset of 𝐻. We set
𝐻𝑛 ∶= span{𝑥𝑘 ∶ 𝑘 = 1, … , 𝑛}. Then 𝐻𝑛 is a nested sequence of finite dimen-
sional subspaces of 𝐻 whose union contains 𝐷 and therefore is dense in 𝐻. We have
dim 𝐻𝑛 ≤ dim 𝐻𝑛+1 , possibly with equality. We inductively construct a basis for span 𝐷
by first choosing a basis of 𝐻1 . Given a basis for 𝐻𝑛 we extend it to a basis of 𝐻𝑛+1
if dim 𝐻𝑛+1 > dim 𝐻𝑛 , otherwise we keep the basis we had. Doing that inductively
from 𝑛 = 1 will give a basis for 𝐻𝑛 for each 𝑛 ∈ ℕ. The union of all these bases is a
countable linearly independent set spanning span 𝐷. Applying the Gram-Schmidt or-
thonormalisation process we can get a countable orthonormal system spanning span 𝐷.
Since span 𝐷 is dense, it follows that 𝐻 has a complete countable orthonormal system.
Using the above theorem we show that there is, up to an isometric isomorphism, there
is only one separable Hilbert space, namely 𝓁2 . Hence 𝓁2 plays the same role as 𝕂𝑁 is
isomorphic to an arbitrary 𝑁-dimensional space.
67
68
Chapter V
Linear Operators
19 Baire’s Theorem
To prove some fundamental properties of linear operators on Banach spaces we need
Baire’s theorem on the intersection of dense open sets.
19.1 Theorem (Baire’s Theorem) Suppose that 𝑋 is a complete metric space and that
⋂
𝑂𝑛 , 𝑛 ∈ ℕ, are open dense subsets of 𝑋. Then 𝑛∈ℕ 𝑂𝑛 is dense in 𝑋.
⋂
Proof. We want to show that for every 𝑥 ∈ 𝑋 and 𝜀 > 0 the ball 𝐵(𝑥, 𝜀) and 𝑛∈ℕ 𝑂𝑛
have non-empty intersection. Fix 𝑥 ∈ 𝑋 and 𝜀 > 0. We show that there exist sequences
(𝑥𝑛 ) in 𝑋 and (𝜀𝑛 ) in ℝ such that 𝑥1 = 𝑥, 𝜀1 = 𝜀, 𝜀𝑛 → 0 and
for all 𝑛 ∈ ℕ. We prove the existence of such sequences inductively. Fix 𝑥 ∈ 𝑋 and
𝜀 > 0. We set 𝑥1 ∶= 𝑥 and 𝜀1 ∶= min{1, 𝜀}. Suppose that we have chosen 𝑥𝑛 and 𝜀𝑛
already. Since 𝑂𝑛 is open and dense in 𝑋 the set 𝐵(𝑥𝑛 , 𝜀𝑛 ) ∩ 𝑂𝑛 is non-empty and open.
Hence we can choose 𝑥𝑛+1 ∈ 𝐵(𝑥𝑛 , 𝜀𝑛 ) and 0 < 𝜀𝑛+1 < min{𝜀𝑛 , 1∕𝑛} such that (19.1)
is true.
From the construction we have 𝐵(𝑥𝑛+1 , 𝜀𝑛+1 ) ⊆ 𝐵(𝑥𝑛 , 𝜀𝑛 ) for all 𝑛 ∈ ℕ and 𝜀𝑛 →
0. Since 𝑋 is complete Cantor’s intersection theorem (Theorem 3.11) implies that
⋂ ⋂
𝑛∈ℕ 𝐵(𝑥𝑛 , 𝜀𝑛 ) is non-empty and so we can choose 𝑦 ∈ 𝑛∈ℕ 𝐵(𝑥𝑛 , 𝜀𝑛 ). By (19.1) we
have
𝑦 ∈ 𝐵(𝑥𝑛+1 , 𝜀𝑛+1 ) ⊆ 𝐵(𝑥𝑛 , 𝜀𝑛 ) ∩ 𝑂𝑛
⋂
for every 𝑛 ∈ ℕ and therefore 𝑦 ∈ 𝑂𝑛 for all 𝑛 ∈ ℕ. Hence 𝑦 ∈ 𝑛∈ℕ 𝑂𝑛 . By
construction 𝐵(𝑥𝑛 , 𝜀𝑛 ) ⊆ 𝐵(𝑥, 𝜀) for all 𝑛 ∈ ℕ and so 𝑦 ∈ 𝐵(𝑥, 𝜀) as well. Since 𝑥 and
⋂
𝜀 > 0 were chosen arbitrarily we conclude that 𝑛∈ℕ 𝑂𝑛 is dense in 𝑋.
69
19.2 Remark In Baire’s Theorem it is essential that the sets 𝑂𝑛 be open in 𝑋 (or satisfy
a suitable other condition). If we set 𝑂1 = ℚ and 𝑂2 = ℝ ⧵ ℚ, then 𝑂1 and 𝑂2 are
dense subsets of ℝ, but 𝑂1 ∩ 𝑂2 = ∅.
19.3 Corollary Suppose that 𝑋 is a non-empty complete metric space and that 𝐶𝑛 ,
⋃
𝑛 ∈ ℕ, is a family of closed sets in 𝑋 with 𝑋 = 𝑛∈ℕ 𝐶𝑛 . Then there exists 𝑛 ∈ ℕ such
that 𝐶𝑛 has non-empty interior.
Proof. If 𝐶𝑛 has empty interior for all 𝑛 ∈ ℕ, then 𝑂𝑛 ∶= 𝑋 ⧵ 𝐶𝑛 is open and dense in
⋂
𝑋 for all 𝑛 ∈ ℕ. By Baire’s Theorem the intersection 𝑛∈ℕ 𝑂𝑛 is dense in 𝑋. Hence
⋃ ⋃ (⋂ )
𝐶𝑛 = (𝑋 ⧵ 𝑂𝑛 ) = 𝑋 ⧵ 𝑂𝑛 ≠ 𝑋.
𝑛∈ℕ 𝑛∈ℕ 𝑛∈ℕ
20.1 Definition (open map) Suppose that 𝑋, 𝑌 are metric spaces. We call a map 𝑓 ∶ 𝑋 →
𝑌 open if 𝑓 maps every open subset of 𝑋 onto an open subset of 𝑌 .
Our aim is to show that linear surjective maps are open. We start by proving a charac-
terisation of open linear maps.
(i) 𝑇 is open;
( )
(ii) There exists 𝑟 > 0 such that 𝐵(0, 𝑟) ⊆ 𝑇 𝐵(0, 1) ;
( )
(iii) There exists 𝑟 > 0 such that 𝐵(0, 𝑟) ⊆ 𝑇 𝐵(0, 1) .
70
Proof.
( We
) prove (i) implies (ii) and hence also (iii). As 𝐵(0, 1) is open, the set
𝑇 𝐵(0, 1) is open in 𝐹 by (i). Since 0 ∈ 𝑇 (𝐵(0, 1)) there exists 𝑟 > 0 such that
( ) ( )
𝐵(0, 𝑟) ⊆ 𝑇 (𝐵(0, 1)) ⊆ 𝑇 𝐵(0, 1) ⊆ 𝑇 𝐵(0, 1) ,
so (iii) follows.
We next prove that (iii) implies (ii). This is the most difficult part of the proof.
Assume that there exists 𝑟 > 0 such that
( )
𝐵(0, 𝑟) ⊆ 𝑇 𝐵(0, 1) .
( )
We show that 𝐵(0, 𝑟∕2) ⊆ 𝑇 𝐵(0, 1) which proves (ii). Hence let 𝑦 ∈ 𝐵(0, 𝑟∕2). Then
( )
2𝑦 ∈ 𝐵(0, 𝑟) and since 𝐵(0, 𝑟) ⊆ 𝑇 𝐵(0, 1) there exists 𝑥1 ∈ 𝐵(0, 1) such that
𝑟
‖2𝑦 − 𝑇 𝑥1 ‖ ≤ .
2
Hence 4𝑦−2𝑇 𝑥1 ∈ 𝐵(0, 𝑟) and by the same argument as before there exists 𝑥2 ∈ 𝐵(0, 1)
such that
𝑟
‖4𝑦 − 2𝑇 𝑥1 − 𝑇 𝑥2 ‖ ≤ .
2
Continuing this way we can construct a sequence (𝑥𝑛 ) in 𝐵(0, 1) such that
𝑟
‖2𝑛 𝑦 − 2𝑛−1 𝑇 𝑥1 − ⋯ − 2𝑇 𝑥𝑛−1 − 𝑇 𝑥𝑛 ‖ ≤
2
for all 𝑛 ∈ ℕ. Dividing by 2𝑛 we get
‖ ∑ 𝑛
‖ 𝑟
‖𝑦 − 2−𝑘 𝑇 𝑥𝑘 ‖ ≤ 𝑛+1
‖ ‖ 2
𝑘=1
∑
∞
∑
∞
2 ‖𝑥𝑘 ‖ ≤
−𝑘
2−𝑘 = 1
𝑘=1 𝑘=1
71
( )
by construction of 𝑥. Hence 𝑦 ∈ 𝑇 𝐵(0, 1) and (ii) follows.
We finally prove that (ii) implies (i). By (ii) and the linearity of 𝑇 we have
( ) ( )
𝑇 𝐵(0, 𝜀) = 𝜀𝑇 𝐵(0, 1)
( )
for every 𝜀 > 0. Because the map 𝑥 → 𝜀𝑥 is a homeomorphism on 𝐹 the set 𝑇 𝐵(0, 𝜀)
is a neighbourhood of zero for every 𝜀 > 0. Let now 𝑈 ⊆ 𝐸 be open and 𝑦 ∈ 𝑇 (𝑈 ).
As 𝑈 is open there exists 𝜀 > 0 such that
𝐵(𝑥, 𝜀) = 𝑥 + 𝐵(0, 𝜀) ⊆ 𝑈 ,
1( )
𝑦= (𝑥 + ℎ) + (−𝑥 + ℎ) ∈ 𝑆̄
2
20.5 Theorem (open mapping theorem) Suppose that 𝐸 and 𝐹 are Banach spaces.
If 𝑇 ∈ (𝐸, 𝐹 ) is surjective, then 𝑇 is open.
Proof. As 𝑇 is surjective we have
⋃ ( )
𝐹 = 𝑇 𝐵(0, 𝑛)
𝑛∈ℕ
( )
with 𝑇 𝐵(0, 𝑛) closed for all 𝑛 ∈ ℕ. Since 𝐹 is complete, by Corollary 19.3 to Baire’s
( )
theorem there exists 𝑛 ∈ ℕ such that 𝑇 𝐵(0, 𝑛) has non-empty interior. Since the map
72
( )
𝑥 → 𝑛𝑥 is a homeomorphism and 𝑇 is linear, the set 𝑇 𝐵(0, 1) has non-empty interior
as well. Now 𝐵(0, 1) is convex and 𝐵(0, 1) = −𝐵(0, 1). The linearity of 𝑇 implies that
( ) ( )
𝑇 𝐵(0, 1) = −𝑇 𝐵(0, 1)
( )
is convex as well. Since we already know that 𝑇 𝐵(0, 1) has non-empty interior,
( )
Lemma 20.4 implies that 𝑇 𝐵(0, 1) is a neighbourhood of zero, that is, there exists
𝑟 > 0 such that
( )
𝐵(0, 𝑟) ⊆ 𝑇 𝐵(0, 1) .
Since 𝐸 is complete Proposition 20.3 implies that 𝑇 is open.
20.6 Corollary (bounded inverse theorem) Suppose that 𝐸 and 𝐹 are Banach spaces
and that 𝑇 ∈ (𝐸, 𝐹 ) is bijective. Then 𝑇 −1 ∈ (𝐹 , 𝐸).
Proof. We know that 𝑇 −1 ∶ 𝐹 → 𝐸 is linear. If 𝑈 ⊆ 𝐸 is open, then by the open
mapping theorem (𝑇 −1 )−1 (𝑈 ) = 𝑇 (𝑈 ) is open in 𝐹 since 𝑇 is surjective. Hence 𝑇 −1 is
continuous by Theorem 5.3, that is, 𝑇 −1 ∈ (𝐹 , 𝐸).
20.7 Corollary Suppose 𝐸 is a Banach space with respect to the norms ‖⋅‖1 and ‖⋅‖2 .
If ‖⋅‖2 is stronger than ‖⋅‖1 , then the two norms are equivalent.
Proof. Let 𝐸1 ∶= (𝐸, ‖⋅‖1 ) and 𝐸2 ∶= (𝐸, ‖⋅‖2 ). Since ‖⋅‖2 is stronger than ‖⋅‖1
the map 𝐼 ∶ 𝑥 → 𝑥 is bounded and bijective from 𝐸2 into 𝐸1 . Since 𝐸1 and 𝐸2 are
Banach spaces the inverse map 𝐼 −1 ∶ 𝑥 → 𝑥 is bounded as well by the bounded inverse
theorem (Corollary 20.6). Hence there exists 𝑐 > 0 such that ‖𝑥‖2 ≤ 𝑐‖𝑥‖1 for all
𝑥 ∈ 𝐸, proving that the two norms are equivalent.
73
21.1 Theorem (closed graph theorem) Suppose that 𝐸 and 𝐹 are Banach spaces and
that 𝑇 ∈ Hom(𝐸, 𝐹 ). Then 𝑇 ∈ (𝐸, 𝐹 ) if and only if 𝑇 has closed graph.
Proof. If 𝑇 ∈ (𝐸, 𝐹 ) and (𝑥𝑛 , 𝑇 𝑥𝑛 ) → (𝑥, 𝑦) then by the continuity 𝑇 𝑥𝑛 → 𝑇 𝑥 = 𝑦,
so (𝑥, 𝑦) = (𝑥, 𝑇 𝑥) ∈ 𝐺(𝑇 ). This shows that 𝐺(𝑇 ) is closed. Now suppose that 𝐺(𝑇 )
is closed. We define linear maps
𝑆 𝑃
← 𝐺(𝑇 ) ←←→
𝐸 ←←→ ← 𝐹, 𝑥 → (𝑥, 𝑇 𝑥) → 𝑇 𝑥.
Then obviously
‖𝑃 (𝑥, 𝑇 𝑥)‖𝐹 = ‖𝑇 𝑥‖𝐹 ≤ ‖(𝑥, 𝑇 𝑥)‖𝑇 .
Hence 𝑃 ∈ (𝐺(𝑇 ), 𝐹 ). Next we note that 𝑆 ∶ 𝐸 → 𝐺(𝑇 ) is an isomorphism because
clearly 𝑆 is surjective and also injective since 𝑆𝑥 = (𝑥, 𝑇 𝑥) = (0, 0) implies that 𝑥 = 0.
Moreover, 𝑆 −1 (𝑥, 𝑇 𝑥) = 𝑥, and therefore
for all 𝑥 ∈ 𝐸. Hence 𝑆 −1 ∈ (𝐺(𝑇 ), 𝐸). Since every closed subspace of a Banach
space is a Banach space, 𝐺(𝑇 ) is a Banach space with respect to the graph norm. By
the bounded inverse theorem (Corollary 20.6) 𝑆 = (𝑆 −1 )−1 ∈ (𝐸, 𝐺(𝑇 )) and so
𝑇 = 𝑃 ◦𝑆 ∈ (𝐸, 𝐹 ) as claimed.
22.1 Theorem (uniform boundedness principle) Suppose that 𝐸 and 𝐹 are Banach
spaces. Let (𝑇𝛼 )𝛼∈𝐴 be a family of linear operators in (𝐸, 𝐹 ) such that sup𝛼∈𝐴 ‖𝑇𝛼 𝑥‖𝐹 <
∞ for all 𝑥 ∈ 𝐸. Then sup𝛼∈𝐴 ‖𝑇𝛼 ‖(𝐸,𝐹 ) < ∞.
Proof. Fix 𝑥0 ∈ 𝐸 and 𝑟 > 0. Then by definition of the operator norm
1 1( )
‖𝑇 ℎ‖ = ‖𝑇 (𝑥0 + ℎ) − 𝑇 (𝑥0 − ℎ)‖ ≤ ‖𝑇 (𝑥0 + ℎ)‖ + ‖𝑇 (𝑥0 − ℎ)‖
2 { 2 }
≤ max ‖𝑇 (𝑥0 + ℎ)‖, ‖𝑇 (𝑥0 − ℎ)‖ ≤ sup ‖𝑇 (𝑥0 + ℎ)‖
‖ℎ‖≤𝑟
for all ℎ ∈ 𝐸 with ‖ℎ‖ < 𝑟. By definition of the operator norm we therefore get
Assume now that sup𝛼∈𝐴 ‖𝑇𝛼 ‖(𝐸,𝐹 ) = ∞. Choose a sequence 𝛼𝑛 ∈ 𝐴 such that
1
‖𝑇𝛼𝑛 ‖(𝐸,𝐹 ) > . (22.2)
4𝑛
74
Let 𝑥0 = 0 and use (22.1) to choose 𝑥1 ∈ 𝐸 such that
1 21
‖𝑥1 − 𝑥0 ‖ < and ‖𝑇𝛼1 𝑥1 ‖ ≥ ‖𝑇 ‖.
3 3 3 𝛼1
Then by (22.1) it is possible to choose 𝑥2 ∈ 𝐸 such that
1 21
‖𝑥2 − 𝑥1 ‖ < and ‖𝑇𝛼2 𝑥2 ‖ ≥ ‖𝑇 ‖.
32 3 32 𝛼2
Continuing that way we inductively choose 𝑥𝑛 ∈ 𝐸 such that
1 21
‖𝑥𝑛 − 𝑥𝑛−1 ‖ < and ‖𝑇𝛼𝑛 𝑥𝑛 ‖ ≥ ‖𝑇 ‖.
3𝑛 3 3𝑛 𝛼𝑛
The sequence (𝑥𝑛 ) is a Cauchy sequence since for 𝑛 > 𝑚 ≥ 0
∑
𝑛
∑𝑛 ( )
1 1 1 1
‖𝑥𝑛 − 𝑥𝑚 ‖ ≤ ‖𝑥𝑘 − 𝑥𝑘−1 ‖ ≤ = − .
𝑘=𝑚+1 𝑘=𝑚+1
3𝑘 2 3𝑚 3𝑛
22.2 Remark There are stronger versions of the uniform boundedness theorem: Either
the family is bounded in the operator norm, or it is pointwise unbounded for a rather
large set of points (see [10, Theorem 5.8] for a more precise statement).
We can apply the above to countable families of linear operators, and in particular to
sequences which converge pointwise. We prove that the limit operator is bounded if it
is the pointwise limit of bounded linear operators between Banach spaces. Note that in
general, the pointwise limit of continuous functions is not continuous.
22.3 Corollary Suppose that 𝐸 and 𝐹 are Banach spaces and that (𝑇𝑛 ) is a sequence
in (𝐸, 𝐹 ) such that 𝑇 𝑥 ∶= lim𝑛→∞ 𝑇𝑛 𝑥 exists for all 𝑥 ∈ 𝐸. Then 𝑇 ∈ (𝐸, 𝐹 ) and
‖𝑇 ‖(𝐸,𝐹 ) ≤ lim inf ‖𝑇𝑛 ‖(𝐸,𝐹 ) < ∞. (22.3)
𝑛→∞
75
Proof. 𝑇 is linear because for 𝑥, 𝑦 ∈ 𝐸 and 𝜆, 𝜇 ∈ 𝕂 we have
Note that in the above proof we cannot in general replace the limit inferior by a limit
because we only know that the sequence ‖𝑇𝑛 ‖ is bounded.
As an application of the uniform boundedness principles we can prove that there
are Fourier series of continuous functions not converging at a given point, say at zero.
22.4 Example Let 𝐶2𝜋 ([−𝜋, 𝜋], ℂ) be the subspace of 𝑢 ∈ 𝐶([−𝜋, 𝜋], ℂ) with 𝑢(−𝜋) =
𝑢(𝜋) equipped with the supremum norm. It represents the space of continuous functions
on ℝ that are 2𝜋-periodic. As 𝐶2𝜋 ([−𝜋, 𝜋], ℂ) ⊆ 𝐿2 ((−𝜋, 𝜋), ℂ) it follows from Exam-
ple 18.4 that the Fourier series of every 𝑓 ∈ 𝐶([−𝜋, 𝜋], ℂ) converges in 𝐿2 ((−𝜋, 𝜋), ℂ).
The question about pointwise convergence is much more delicate. We use the uniform
boundedness principle to show that there are continuous functions, where the Fourier
series diverges. To do so let 𝑢 ∈ 𝐶([−𝜋, 𝜋], ℂ) with Fourier series
1 ∑
∞ 𝜋
𝑢(𝑥) = 𝑢(𝑡)𝑒−𝑖𝑘𝑡 𝑑𝑡𝑒𝑖𝑘𝑥
2𝜋 𝑘=−∞ ∫−𝜋
We will show that the operators 𝑇𝑛 ∶ 𝐶2𝜋 ([−𝜋, 𝜋], ℂ) → ℂ given by the partial sum
1 ∑
𝑛 𝜋
𝑇𝑛 𝑢 ∶= 𝑢𝑛 (0) ∶= 𝑢(𝑡)𝑒−𝑖𝑘𝑡 𝑑𝑡 (22.4)
2𝜋 𝑘=−𝑛 ∫−𝜋
of the Fourier series evaluated at 𝑥 = 0 is a bounded linear operator. It turns out that
‖𝑇𝑛 ‖ → ∞ as 𝑛 → ∞, so the family of operators (𝑇𝑛 )𝑛∈ℕ is not bounded. Hence,
by the uniform boundedness principle there must exist 𝑢 ∈ 𝐶2𝜋 ([−𝜋, 𝜋], ℂ) so that
𝑇𝑛 𝑢 = 𝑢𝑛 (0) is not bounded, that is, the Fourier series of 𝑢 does not converge at 𝑥 = 0.
One can replace 𝑥 = 0 by an arbitrary 𝑥 ∈ [−𝜋, 𝜋]. One can also use Baire’s Theorem
to show that for most 𝑢 ∈ 𝐶2𝜋 ([−𝜋, 𝜋], ℂ), the Fourier series diverges for instance on
[−𝜋, 𝜋] ∩ ℚ. We do not do that here, but refer to [10, Section 5.11].
We now show determine ‖𝑇𝑛 ‖ and show that ‖𝑇𝑛 ‖ → ∞ as 𝑛 → ∞. To do so we
set
∑𝑛
𝐷𝑛 (𝑡) ∶= 𝑒𝑖𝑘𝑡 .
𝑘=−𝑛
76
We can then write the partial sum of the Fourier series of 𝑢 in the form
∑
𝑛
1 ∑ 𝑖𝑘𝑥
𝑛 𝜋
𝑢𝑛 (𝑥) = 𝑐𝑘 𝑒𝑖𝑘𝑥 = 𝑒 𝑢(𝑡)𝑒−𝑖𝑘𝑡 𝑑𝑡
𝑘=−𝑛
2𝜋 𝑘=−𝑛 ∫−𝜋
1
𝜋 ∑ 𝑛
1
𝜋
= 𝑢(𝑡) 𝑒𝑖𝑘((𝑥−𝑡) 𝑑𝑡 = 𝑢(𝑡)𝐷𝑛 (𝑥 − 𝑡) 𝑑𝑡.
2𝜋 ∫−𝜋 𝑘=−𝑛
2𝜋 ∫−𝜋
Using the formula for the partial sum of a geometric series we also see that
∑
𝑛
∑
2𝑛
𝑖𝑘𝑡 −𝑖𝑛𝑡
𝐷𝑛 (𝑡) = 𝑒 =𝑒 (𝑒𝑖𝑡 )𝑘
𝑘=−𝑛 𝑘=0
𝑒𝑖(2𝑛+1)𝑡 − 1 𝑒𝑖(𝑛+1∕2)𝑡 − 𝑒−𝑖(𝑛+1∕2)𝑡 sin(𝑛 + 1∕2)𝑡
= 𝑒−𝑖𝑛𝑡 = = .
𝑒𝑖𝑡 − 1 𝑒𝑖𝑡∕2 − 𝑒−𝑖𝑡∕2 sin(𝑡∕2)
sin(𝑛 + 1∕2)𝑡
𝐷𝑛 (𝑡) = (22.5)
sin(𝑡∕2)
𝑦
2𝑛 + 1
−𝜋 0 𝜋 𝑡
77
are bounded since, using (22.6).
1
|𝑇𝑛 𝑢| ≤ ‖𝐷 ‖ ‖𝑢‖∞ .
2𝜋 𝑛 1
Hence,
1
‖𝐷 ‖ .
‖𝑇𝑛 ‖ ≤
2𝜋 𝑛 1
We would like to show equality. If we set 𝑢 ∶= sign 𝐷𝑛 , then 𝑢(𝑡)𝐷𝑛 (𝑡) = |𝐷𝑛 (𝑡)| and
thus 𝜋
1 1
|𝑇𝑛 𝑢| = |𝐷 (𝑡)| 𝑑𝑡 = ‖𝐷 ‖ .
2𝜋 ∫−𝜋 𝑛 2𝜋 𝑛 1
Unfortunately, 𝑢 = sign 𝐷𝑛 is not continous, otherwise the above would imply that
1
‖𝑇𝑛 ‖ ≥
‖𝐷 ‖ (22.8)
2𝜋 𝑛 1
However, we can approximate 𝑢 pointwise by functions 𝑢𝑘 ∈ 𝐶2𝜋 ([−𝜋, 𝜋], ℝ) with
values in [−1, 1] by putting in a line at every jump of with slope ±𝑘 as indicated in
Figure 22.2. The dominated convergence theorem then implies that
𝑦 sign 𝐷𝑛
𝑢𝑘
−𝜋 0 𝜋 𝑡
𝜋 𝜋 𝜋
lim 𝑢𝑘 (𝑡)𝐷𝑛 (𝑡) 𝑑𝑡 = sign(𝐷𝑛 (𝑡))𝐷𝑛 (𝑡) 𝑑𝑡 = |𝐷𝑛 (𝑡)| 𝑑𝑡 = ‖𝐷𝑛 ‖1
𝑘→∞ ∫−𝜋 ∫−𝜋 ∫−𝜋
and hence (22.8) is valid. In particular 2𝜋‖𝑇𝑛 ‖ = ‖𝐷𝑛 ‖1 for all 𝑛 ∈ ℕ. We finally need
to show that ‖𝑇𝑛 ‖ → ∞ or equivalently that ‖𝐷𝑛 ‖1 → ∞ as 𝑛 → ∞. As sin(𝑡∕2) ≤
|𝑡|∕2 for all 𝑡 ∈ ℝ we first note that
𝜋 𝜋 𝜋
| sin(𝑛 + 1∕2)𝑡|
|𝐷𝑛 (𝑡)| 𝑑𝑡 = 2 |𝐷𝑛 (𝑡)| 𝑑𝑡 = 2 𝑑𝑡
∫−𝜋 ∫0 ∫0 | sin(𝑡∕2)|
𝜋
| sin(𝑛 + 1∕2)𝑡|
≥4 𝑑𝑡 =∶ 4𝐼𝑛
∫0 𝑡
for all 𝑛 ∈ ℕ. Using the substitution 𝑠 = (𝑛 + 1∕2)𝑡 we see that
𝜋
| sin(𝑛 + 1∕2)𝑡| (𝑛+1∕2)𝜋
| sin 𝑠| ∑ 𝑘𝜋 | sin 𝑠|
𝑛
𝐼𝑛 = 𝑑𝑡 = 𝑑𝑠 ≥ 𝑑𝑠
∫0 𝑡 ∫0 𝑠 ∫
𝑘=1 (𝑘−1)𝜋
𝑠
∑𝑛
1
𝑘𝜋
1 ∑ 1 𝑛→∞
𝑛
≥ | sin 𝑠| 𝑑𝑠 = ← ∞
←←←←←←←←→
𝑘=1
𝑘𝜋 ∫(𝑘−1)𝜋 𝜋 1 𝑘
because of the divergence of the harmonic series. Hence, ‖𝐷𝑛 ‖1 → ∞ as claimed.
78
23 Closed Operators
In this section we look at a class of operators which are not necessarily bounded, but
still sharing many properties with bounded operators.
23.2 Remark (a) The graph of a linear operator 𝐴 ∶ 𝐷(𝐴) ⊆ 𝐸 → 𝐹 is closed if and
only if 𝑥𝑛 → 𝑥 in 𝐸 and 𝐴𝑥𝑛 → 𝑦 in 𝐹 imply that 𝑥 ∈ 𝐷(𝐴) and 𝐴𝑥 = 𝑦. The condition
looks similar to continuity. The difference is that continuity means that 𝑥𝑛 → 𝑥 in 𝐸
implies that the sequence (𝐴𝑥𝑛 ) automatically converges, whereas this is assumed in
case of a closed operator.
(b) Note that 𝐴 ∶ 𝐷(𝐴) → 𝐹 is always continuous if we consider 𝐷(𝐴) with the
graph norm ‖𝑥‖𝐴 ∶= ‖𝑥‖𝐸 + ‖𝐴𝑥‖𝐹 .
Classical examples of closed but unbounded operators are differential operators. We
demonstrate this with the simplest possible example.
23.3 Example Consider the vector space of continuous functions 𝐸 = 𝐶([0, 1]). De-
fine the operator 𝐴𝑢 ∶= 𝑢′ with domain
23.4 Theorem Suppose that 𝐸 and 𝐹 are Banach spaces and that 𝐴 ∶ 𝐷(𝐴) ⊆ 𝐸 → 𝐹
is an injective closed operator.
(i) Then 𝐴−1 ∶ im(𝐴) ⊆ 𝐹 → 𝐸 is a closed operator with domain 𝐷(𝐴−1 ) = im(𝐴).
79
Proof. (i) Note that
We can also characterise closed operators in terms of properties of the graph norm on
𝐷(𝐴) as introduced in Definition 9.5.
23.5 Proposition Suppose that 𝐸 and 𝐹 are Banach spaces. Then, the operator 𝐴 ∶ 𝐷(𝐴) ⊆
𝐸 → 𝐹 is closed if and only if 𝐷(𝐴) is a Banach space with respect to the graph norm
‖𝑥‖𝐴 ∶= ‖𝑥‖𝐸 + ‖𝐴𝑥‖𝐹 .
Proof. We first assume that 𝐴 is closed and show that 𝐷(𝐴) is complete with respect
to the graph norm. Let (𝑥𝑛 ) be a Cauchy sequence in (𝐷(𝐴), ‖⋅‖𝐴 ). Fix 𝜀 > 0. Then
there exists 𝑛0 ∈ ℕ such that
for all 𝑚, 𝑛 > 𝑛0 . Hence (𝑥𝑛 ) and (𝐴𝑥𝑛 ) are Cauchy sequences in 𝐸 and 𝐹 , respectively.
By the completeness of 𝐸 and 𝐹 we conclude that 𝑥𝑛 → 𝑥 and 𝐴𝑥𝑛 → 𝑦. Since 𝐴 is
closed we have 𝑥 ∈ 𝐷(𝐴) and 𝐴𝑥 = 𝑦, so
implies that (𝑥𝑛 ) is a Cauchy sequence with respect to the graph norm in 𝐷(𝐴). By the
completeness of 𝐷(𝐴) with respect to the graph norm we have 𝑥𝑛 → 𝑥 in the graph
norm. By the definition of the graph norm this in particular implies 𝑥 ∈ 𝐷(𝐴) and
𝐴𝑥𝑛 → 𝐴𝑥, so 𝐴𝑥 = 𝑦. Hence 𝐴 is closed.
In the following corollary we look at the relationship between continuous and closed
operators. Continuous means with respect to the norm on 𝐷(𝐴) induced by the norm
in 𝐸.
80
23.6 Corollary Suppose that 𝐸, 𝐹 are Banach spaces and that 𝐴 ∶ 𝐷(𝐴) ⊆ 𝐸 → 𝐹 is
a linear operator. Then 𝐴 ∈ (𝐷(𝐴), 𝐹 ) is closed if and only if 𝐷(𝐴) is closed in 𝐸.
Proof. Suppose that 𝐴 ∈ (𝐷(𝐴), 𝐹 ). By Proposition 9.7, the graph norm on 𝐷(𝐴)
is equivalent to the norm of 𝐸 restricted to 𝐷(𝐴). Since 𝐴 is closed we know from
Proposition 23.5 that 𝐷(𝐴) is complete with respect to the graph norm, and therefore
with respect to the norm in 𝐸. Hence 𝐷(𝐴) is closed in 𝐸. Suppose now that 𝐷(𝐴)
is closed in 𝐸. Since 𝐸 is a Banach space, also 𝐷(𝐴) is a Banach space. Hence, the
closed graph theorem (Theorem 21.1) implies that 𝐴 ∈ (𝐷(𝐴), 𝐹 ).
As a consequence of the above result the domain of a closed and unbounded operator
𝐴 ∶ 𝐷(𝐴) ⊆ 𝐸 → 𝐹 is never a closed subset of 𝐸.
(i) 𝐴 is closable.
̄ ∶= {𝑥 ∈ 𝐸 ∶ (𝑥, 𝑦) ∈ 𝐺(𝐴)}
𝐷(𝐴)
81
and its graph by 𝐺(𝐴) ̄ = 𝐺(𝐴). If 𝑥, 𝑧 ∈ 𝐷(𝐴),̄ then by the above construction there
exist sequences (𝑥𝑛 ) and (𝑧𝑛 ) in 𝐷(𝐴) with 𝐴𝑥𝑛 → 𝐴𝑥
̄ and 𝐴𝑧𝑛 → 𝐴𝑧.̄ Hence
̄ + 𝜇 𝐴𝑧
𝐴(𝜆𝑥𝑛 + 𝜇𝑧𝑛 ) = 𝜆𝐴𝑥𝑛 + 𝜇𝐴𝑧𝑛 → 𝜆𝐴𝑥 ̄
Since 𝜆𝑥𝑛 + 𝜇𝑧𝑛 → 𝜆𝑥 + 𝜇𝑧, by definition of 𝐴̄ we get (𝜆𝑥 + 𝜇𝑧, 𝜆𝐴𝑥 ̄ + 𝜇𝐴𝑧)
̄ ∈ 𝐺(𝐴) ̄
and so 𝐴(𝜆𝑥 + 𝜇𝑧) = 𝜆𝐴𝑥 + 𝜇𝐴𝑧. Hence 𝐷(𝐴) is a subspace of 𝐸 and 𝐴 is linear.
̄ ̄ ̄ ̄ ̄
Because 𝐺(𝐴) ̄ = 𝐺(𝐴) is closed, it follows that 𝐴̄ is a closed operator. Moreover, 𝐴̄ is
the closure of 𝐴 because 𝐺(𝐴) must be contained in the graph of any closed operator
extending 𝐴.
We can use the above in particular to construct extensions of bounded linear operators
defined on dense subspaces.
24.3 Corollary (Extension of linear operators) Suppose that 𝐸, 𝐹 are Banach spaces,
and that 𝐸0 is a dense subspace of 𝐸. If 𝑇0 ∶ (𝐸0 , 𝐹 ), then there exists a unique op-
erator 𝑇 ∈ (𝐸, 𝐹 ) such that 𝑇 𝑥 = 𝑇0 𝑥 for all 𝑥 ∈ 𝐸0 . Moreover, ‖𝑇0 ‖(𝐸0 ,𝐹 ) =
‖𝑇 ‖(𝐸,𝐹 ) .
Hence ‖𝑇 ‖(𝐸,𝐹 ) ≤ ‖𝑇0 ‖(𝐸0 ,𝐹 ) . Since clearly ‖𝑇 ‖(𝐸,𝐹 ) ≥ ‖𝑇0 ‖(𝐸0 ,𝐹 ) , the equality of
the operator norms follow. Now Corollary 23.6 implies that 𝐷(𝑇 ) is closed in 𝐸, and
since 𝐸0 ⊆ 𝐷(𝑇 ) is dense in 𝐸 we get 𝐷(𝑇 ) = 𝐸.
and
𝐶𝑐∞ (Ω) ∶= {𝑢 ∈ 𝐶 ∞ (Ω) ∶ supp(𝑢) ⊆ Ω is compact}.
The elements of 𝐶𝑐∞ (Ω) are called test functions.
82
24.5 Example Let 𝐼 = [𝑎, 𝑏] be a compact interval, 𝑝 ∈ 𝐶 1 (𝐼), 𝑞 ∈ 𝐶(𝐼) with 𝑝(𝑥) > 0
for all 𝑥 ∈ 𝐼 and 𝛼𝑖 , 𝛽𝑖 ∈ ℝ with 𝛼𝑖2 + 𝛽𝑖2 ≠ 0 (𝑖 = 1, 2). We want to reformulate the
Sturm-Liouville boundary value problem
−(𝑝𝑢′ )′ + 𝑞𝑢 = 𝑓 in (𝑎, 𝑏),
𝛼1 𝑢(𝑎) − 𝛽1 𝑢′ (𝑎) = 0 (24.1)
𝛼2 𝑢(𝑎) + 𝛽2 𝑢′ (𝑎) = 0
in a Hilbert space setting as a problem in 𝐿2 ((𝑎, 𝑏), ℝ). Functions in 𝐿2 ((𝑎, 𝑏), ℝ) are
not generally differentiable, so we first define a linear operator on a smaller domain by
setting
𝐴0 ∶= −(𝑝𝑢′ )′ + 𝑞𝑢
for all 𝑢 ∈ 𝐷(𝐴0 ), where
𝐷(𝐴0 ) ∶= {𝑢 ∈ 𝐶 ∞ (𝐼) ∶ 𝛼1 𝑢(𝑎) − 𝛽1 𝑢′ (𝑎) = 𝛼2 𝑢(𝑎) + 𝛽2 𝑢′ (𝑎) = 0}.
The idea is to deal with the boundary conditions by incorporating them into the defini-
tion of the domain of the differential operator. We note that if 𝐴0 𝑢 = 𝑓 with 𝑢 ∈ 𝐷(𝐴0 ),
then 𝑢 is a solution of (24.1).
We would like to deal with a closed operator, but unfortunately, 𝐴0 is not closed.
We will show that 𝐴0 is closable as an operator in 𝐿2 ((𝑎, 𝑏), ℝ). Clearly 𝐷(𝐴0 ) ⊆
𝐿2 ((𝑎, 𝑏), ℝ). Moreover, 𝐶𝑐∞ ((𝑎, 𝑏)) ⊆ 𝐷(𝐴0 ). Using integration by parts twice we see
that
𝑏 𝑏 𝑏
|𝑏
𝜑𝐴0 𝑢 𝑑𝑥 = 𝑢𝐴0 𝜑 𝑑𝑥 + 𝑝(𝑢𝜑′ − 𝜑𝑢′ )| = 𝑢𝐴0 𝜑 𝑑𝑥 (24.2)
∫𝑎 ∫𝑎 | 𝑎 ∫𝑎
for all 𝑢 ∈ 𝐷(𝐴0 ) and 𝜑 ∈ 𝐶𝑐∞ ((𝑎, 𝑏)) because 𝜑(𝑎) = 𝜑′ (𝑎) = 𝜑(𝑏) = 𝜑′ (𝑏) = 0.
Suppose now that 𝑢𝑛 ∈ 𝐷(𝐴0 ) such that 𝑢𝑛 → 0 in 𝐿2 ((𝑎, 𝑏), ℝ) and 𝐴0 𝑢𝑛 → 𝑣 in
𝐿2 ((𝑎, 𝑏), ℝ). Using (24.2) we and the Cauchy-Schwarz inequality we get
𝑏 𝑏
| | | |
| 𝜑𝐴0 𝑢𝑛 𝑑𝑥| = | 𝑢 𝐴 𝜑 𝑑𝑥| ≤ ‖𝑢𝑛 ‖2 ‖𝐴0 𝜑‖2 → 0
|∫𝑎 | |∫𝑎 𝑛 0 |
as 𝑛 → ∞. Therefore,
𝑏 𝑏
𝜑𝐴0 𝑢𝑛 𝑑𝑥 → 𝜑𝑣 𝑑𝑥 = 0
∫𝑎 ∫𝑎
for all 𝜑 ∈ 𝐶𝑐∞ ((𝑎, 𝑏)). One can show that 𝐶𝑐∞ ((𝑎, 𝑏)) is dense in 𝐿2 ((𝑎, 𝑏), ℝ), so
that 𝑣 = 0 in 𝐿2 ((𝑎, 𝑏), ℝ). By Theorem 24.2 the operator 𝐴0 is closable. Denote
its closure by 𝐴 with domain 𝐷(𝐴). Since 𝐶𝑐∞ ((𝑎, 𝑏)) is dense in 𝐿2 ((𝑎, 𝑏), ℝ) and
𝐶𝑐∞ ((𝑎, 𝑏)) ⊆ 𝐷(𝐴0 ) it follows that 𝐷(𝐴) is a dense subset of 𝐿2 ((𝑎, 𝑏), ℝ).
Instead of (24.1) it is common to study the abstract equation
𝐴𝑢 = 𝑓
in 𝐿2 ((𝑎, 𝑏), ℝ). The solutions may not be differentiable, but if they are, then they are
solutions to (24.1). We call solutions of the abstract equation generalised solutions of
(24.1). The advantage of that setting is that we can use all the Hilbert space theory
including Fourier series expansions of solutions.
83
84
Chapter VI
Duality
25 Dual Spaces
In this section we discuss the space of linear operators from a normed space over 𝕂 into
𝕂. This is called the “dual space” and plays an important role in many applications.
We give some precise definition and then examples.
25.1 Definition (Dual space) If 𝐸 is a normed vector space over 𝕂, then we call
𝐸 ′ ∶= (𝐸, 𝕂)
the dual space of 𝐸. The elements of 𝐸 ′ are often referred to as bounded linear func-
tionals on 𝐸. The operator norm on (𝐸, 𝕂) is called the dual norm on 𝐸 ′ and is
denoted by ‖⋅‖𝐸 ′ . If 𝑓 ∈ 𝐸 ′ , then we define
⟨𝑓 , 𝑥⟩ ∶= 𝑓 (𝑥).
𝐸 ′ → 𝕂, 𝑓 → ⟨𝑓 , 𝑥⟩
and as a consequence
|⟨𝑓 , 𝑥⟩| ≤ ‖𝑓 ‖𝐸 ′ ‖𝑥‖𝐸
for all 𝑓 ∈ 𝐸 ′ and 𝑥 ∈ 𝐸. Hence the duality map ⟨⋅ , ⋅⟩ ∶ 𝐸 ′ × 𝐸 → 𝕂 is continuous.
85
We next give some examples of dual spaces. We also look at specific representations
of dual spaces and the associated duality map, which is useful in many applications.
These representations are often not easy to get.
25.3 Example Let 𝐸 = 𝕂𝑁 . Then the dual space is given by all linear maps from 𝕂𝑁
to 𝕂. Such maps can be represented by a 1 × 𝑁 matrix. Hence the elements of 𝕂𝑁
are column vectors, and the elements of (𝕂𝑁 )′ are row vectors, and the duality map is
simply matrix multiplication. More precisely, if
⎡ 𝑥1 ⎤
𝑓 = [𝑓1 , … , 𝑓𝑁 ] ∈ (𝕂𝑁 )′ and 𝑥 = ⎢ ⋮ ⎥ ∈ 𝕂𝑁 ,
⎢ ⎥
⎣𝑥𝑁 ⎦
then
⎡ 𝑥1 ⎤ ∑𝑁
⟨𝑓 , 𝑥⟩ = [𝑓1 , … , 𝑓𝑁 ] ⎢ ⋮ ⎥ = 𝑓 𝑥 ,
⎢ ⎥ 𝑘=1 𝑘 𝑘
⎣𝑥𝑁 ⎦
which is almost like the dot product except for the complex conjugate in the second
argument. By identifying the row vector 𝑓 with the column vector 𝑓 𝑇 we can identify
the dual of 𝕂𝑁 with itself.
Note however, that the dual norm depends by definition on the norm of the original
space. We want to compute the dual norm for the above identification by looking at
𝐸𝑝 = 𝕂𝑁 with the 𝑝-norm as defined in (7.1). We are going to show that, with the
above identification, ( )
1 1
𝐸𝑝′ = 𝐸𝑝′ + ′ =1
𝑝 𝑝
with equal norms. By Hölder’s inequality (Proposition 7.2)
|⟨𝑓 , 𝑥⟩| ≤ |𝑓 |𝑝′ |𝑥|𝑝 , (25.1)
for all 𝑥 ∈ 𝕂𝑁 and 1 ≤ 𝑝 ≤ ∞. Hence by definition of the dual norm ‖𝑓 ‖𝐸𝑝′ ≤ |𝑓 |𝑝′ .
We show that there is equality for 1 ≤ 𝑝 ≤ ∞, that is,
‖𝑓 ‖𝐸𝑝′ = |𝑓 |𝑝′ . (25.2)
𝑓𝑘 𝑥𝑘 =
if 𝑝 = ∞,
′
𝑓𝑘 𝜆𝑘 = |𝑓𝑘 | = |𝑓𝑘 |𝑝
86
′
Now if 𝑝 = ∞ then |𝑓 |𝑝𝑝′ −1 = |𝑓 |01 = 1, so ‖𝑓 ‖𝐸𝑝′ = |𝑓 |1 . If 𝑝 ∈ (1, ∞) then, by
definition of 𝑥 and since 𝑝′ = 𝑝∕(𝑝 − 1), we have
′
|𝑓𝑘 |𝑝 −1 = |𝑓𝑘 |1∕(𝑝−1) = |𝑥𝑘 |
and so
′
(∑
𝑁 )1−1∕𝑝′ (∑
𝑁 )1∕𝑝
|𝑓 |𝑝𝑝′ −1 = |𝑓𝑘 |𝑝∕(𝑝−1)
= |𝑥𝑘 |𝑝 = |𝑥|𝑝 .
𝑘=1 𝑘=1
so ‖𝑓 ‖𝐸𝑝′ = |𝑓 |∞ if 𝑝 = 1 as well.
We next look at an infinite dimensional analogue to the above example, namely the
sequence spaces 𝓁𝑝 from Section 7.2.
for all 𝑥 ∈ 𝓁𝑝 and 𝑦 ∈ 𝓁𝑝′ , where and 1 ≤ 𝑝 ≤ ∞. This means that for every fixed
𝑦 ∈ 𝓁𝑝′ , the linear functional 𝑥 → ⟨𝑦, 𝑥⟩ is bounded, and therefore an element of the
dual space (𝓁𝑝 )′ . Hence, if we identify 𝑦 ∈ 𝓁𝑝′ with the map 𝑥 → ⟨𝑦, 𝑥⟩, then we
naturally have
𝓁𝑝′ ⊆ 𝓁𝑝′
for 1 ≤ 𝑝 ≤ ∞. The above identification is essentially the same as the one in the
previous example on finite dimensional spaces. We prove now that
(i) 𝓁𝑝′ = 𝓁𝑝′ if 1 ≤ 𝑝 < ∞ with equal norm;
𝑦𝑘 ∶= ⟨𝑓 , 𝑒𝑘 ⟩
∑
𝑛
⟨𝑓𝑛 , 𝑥⟩ ∶= 𝑦𝑘 𝑥𝑘
𝑘=0
87
∑𝑛
If we set 𝑥(𝑛) ∶= 𝑘=0
𝑥𝑘 𝑒𝑘 , then
⟨ ∑ 𝑛 ⟩ ∑𝑛
∑
𝑛
⟨𝑓 , 𝑥 ⟩ = 𝑓 ,
(𝑛)
𝑥𝑘 𝑒𝑘 = 𝑥𝑘 ⟨𝑓 , 𝑒𝑘 ⟩ = 𝑦𝑘 𝑥𝑘 = ⟨𝑓𝑛 , 𝑥⟩ (25.3)
𝑘=0 𝑘=0 𝑘=0
∑
∞
𝑥𝑘 𝑦𝑘 = lim ⟨𝑓𝑛 , 𝑥⟩ = lim ⟨𝑓 , 𝑥(𝑛) ⟩ = ⟨𝑓 , 𝑥⟩,
𝑛→∞ 𝑛→∞
𝑘=0
The first equality holds since |(𝑦1 , … , 𝑦𝑛 )|𝑝′ defines a monotone increasing sequence
with limit |𝑦|𝑝′ . In particular we conclude that 𝑦 ∈ 𝓁𝑝′ .
We next show that |𝑦|𝑝′ = ‖𝑓 ‖𝓁𝑝′ . By Hölder’s inequality |⟨𝑓 , 𝑥⟩| = |⟨𝑦, 𝑥⟩| ≤
|𝑦|𝑝′ |𝑥|𝑝 , and so ‖𝑓 ‖𝓁𝑝′ ≤ |𝑦|𝑝′ . To prove the reverse inequality choose 𝑥(𝑛) ∈ 𝓁𝑝
such that ⟨𝑓 , 𝑥(𝑛) ⟩ = |(𝑦0 , … , 𝑦𝑛 )|𝑝′ |𝑥(𝑛) |𝑝 , which is possible by the construction in the
previous example. Thus by definition of the dual norm, |(𝑦0 , … , 𝑦𝑛 )|𝑝′ ≤ ‖𝑓 ‖𝓁𝑝′ for all
𝑛 ∈ ℕ and hence,
|𝑦|𝑝′ = sup |(𝑦0 , … , 𝑦𝑛 )|𝑝′ ≤ ‖𝑓 ‖𝓁𝑝′
𝑛∈ℕ
𝑔→ 𝑓 (𝑥)𝑔(𝑥) 𝑑𝑥,
∫Ω
88
( )′ ( )′
then we have 𝑓 ∈ 𝐿𝑝 (Ω) . With that identification 𝐿𝑝′ (Ω) ⊆ 𝐿𝑝 (Ω) for 1 ≤ 𝑝 ≤
∞. Using the Radon-Nikodym Theorem from measure theory one can show that
( )′
𝐿𝑝 (Ω) = 𝐿𝑝′ (Ω)
if 1 ≤ 𝑝 < ∞ and ( )′
𝐿1 (Ω) ⊂ 𝐿∞ (Ω)
with proper inclusion.
Let 𝑀 be a subspace and 𝑓0 ∈ Hom(𝑀, ℝ) with 𝑓0 (𝑥) ≤ 𝑝(𝑥) for all 𝑥 ∈ 𝑀. Then
there exists an extension 𝑓 ∈ Hom(𝐸, ℝ) such that 𝑓 |𝑀 = 𝑓0 and 𝑓 (𝑥) ≤ 𝑝(𝑥) for all
𝑥 ∈ 𝐸.
Proof. (a) We first show that 𝑓0 can be extended as required if 𝑀 has co-dimension
one. More precisely, let 𝑥0 ∈ 𝐸 ⧵𝑀 and assume that span(𝑀 ∪{𝑥0 }) = 𝐸. As 𝑥0 ∉ 𝑀
we can write every 𝑥 ∈ 𝐸 in the form
𝑥 = 𝑚 + 𝛼𝑥0
𝑓𝑐 (𝑚 + 𝛼𝑥) ∶= 𝑓0 (𝑚) + 𝑐𝛼
is well defined, and 𝑓𝑐 (𝑚) = 𝑓0 (𝑚) for all 𝑚 ∈ 𝑀. We want to show that it is possible
to choose 𝑐 ∈ ℝ such that 𝑓𝑐 (𝑥) ≤ 𝑝(𝑥) for all 𝑥 ∈ 𝐸. Equivalently
for all 𝑚 ∈ 𝑀 and 𝛼 ∈ ℝ. By the positive homogeneity of 𝑝 and the linearity of 𝑓 the
above is equivalent to
89
Hence we need to show that we can choose 𝑐 such that
𝑐 ≤ 𝑝(𝑥0 + 𝑚∕𝛼) − 𝑓0 (𝑚∕𝛼)
𝑐 ≥ −𝑝(−𝑥0 − 𝑚∕𝛼) + 𝑓0 (−𝑚∕𝛼)
for all 𝑚 ∈ 𝑀 and 𝛼 ∈ ℝ. Note that ±𝑚∕𝛼 is just an arbitrary element of 𝑀, so the
above conditions reduce to
𝑐 ≤ 𝑝(𝑥0 + 𝑚) − 𝑓0 (𝑚)
𝑐 ≥ −𝑝(−𝑥0 + 𝑚) + 𝑓0 (𝑚)
for all 𝑚1 , 𝑚2 ∈ 𝑀. Using the sub-additivity of 𝑝 we can verify this condition since
We next want to get a version of the Hahn-Banach theorem for vector spaces over ℂ.
To do so we need to strengthen the assumptions on 𝑝.
26.3 Lemma Let 𝑝 be a semi-norm on 𝐸. Then 𝑝(𝑥) ≥ 0 for all 𝑥 ∈ 𝐸 and 𝑝(0) = 0.
Moreover, |𝑝(𝑥) − 𝑝(𝑦)| ≤ 𝑝(𝑥 − 𝑦) for all 𝑥, 𝑦 ∈ 𝐸 (reversed triangle inequality).
Proof. Firstly by (ii) we have 𝑝(0) = 𝑝(0𝑥) = 0𝑝(𝑥) = 0 for all 𝑥 ∈ 𝐸, so 𝑝(0) = 0. If
𝑥, 𝑦 ∈ 𝐸, then by (i)
90
so that
𝑝(𝑥) − 𝑝(𝑦) ≤ 𝑝(𝑥 − 𝑦).
Interchanging the roles of 𝑥 and 𝑦 and using (ii) we also have
Combining the two inequalities we get |𝑝(𝑥) − 𝑝(𝑦)| ≤ 𝑝(𝑥 − 𝑦) for all 𝑥, 𝑦 ∈ 𝐸.
Choosing 𝑦 = 0 we get 0 ≤ |𝑝(𝑥)| ≤ 𝑝(𝑥), so 𝑝(𝑥) ≥ 0 for all 𝑥 ∈ 𝐸.
We can now prove a version of the Hahn-Banach Theorem for vector spaces over ℂ.
for all 𝑥 ∈ 𝑀. In particular, ℎ0 (𝑥) = −𝑔0 (𝑖𝑥) and so 𝑔0 (𝑖𝑥) + ℎ0 (𝑥) = 0 for all 𝑥 ∈ 𝑀.
Therefore,
𝑓0 (𝑥) = 𝑔0 (𝑥) − 𝑖𝑔0 (𝑖𝑥) (26.1)
for all 𝑥 ∈ 𝑀. We now consider 𝐸 as a vector space over ℝ. We denote that vector
space by 𝐸ℝ . We can consider 𝑀ℝ as a vector subspace of 𝐸ℝ . Now clearly 𝑔0 ∈
Hom(𝑀ℝ , ℝ) and by assumption
for all 𝑥 ∈ 𝑀ℝ . By Theorem 26.1 there exists 𝑔 ∈ Hom(𝐸ℝ , ℝ) such that 𝑔|𝑀ℝ = 𝑔0
and
𝑔(𝑥) ≤ 𝑝(𝑥)
for all 𝑥 ∈ 𝐸ℝ . We set
𝑓 (𝑥) ∶= 𝑔(𝑥) − 𝑖𝑔(𝑖𝑥)
for all 𝑥 ∈ 𝐸ℝ , so that 𝑓 |𝑀 = 𝑓0 by (26.1). To show that 𝑓 is linear from 𝐸 to ℂ we
only need to look at multiplication by 𝑖 because 𝑓 is linear over ℝ. We have
91
for all 𝑥 ∈ 𝐸. It remains to show that |𝑓 (𝑥)| ≤ 𝑝(𝑥). For fixed 𝑥 ∈ 𝐸 we choose 𝜆 ∈ ℂ
with |𝜆| = 1 such that 𝜆𝑓 (𝑥) = |𝑓 (𝑥)|. Then since |𝑓 (𝑥)| ∈ ℝ and using the definition
of 𝑓
|𝑓 (𝑥)| = 𝜆𝑓 (𝑥) = 𝑓 (𝜆𝑥) = 𝑔(𝜆𝑥) ≤ 𝑝(𝜆𝑥) = |𝜆|𝑝(𝑥) = 𝑝(𝑥).
This completes the proof of the theorem.
From the above version of the Hahn-Banach theorem we can prove the existence of
bounded linear functionals with various special properties.
26.5 Corollary Suppose that 𝐸 is a normed space. Then for every 𝑥 ∈ 𝐸, 𝑥 ≠ 0, there
exists 𝑓 ∈ 𝐸 ′ such that ‖𝑓 ‖𝐸 ′ = 1 and ⟨𝑓 , 𝑥⟩ = ‖𝑥‖𝐸 .
Proof. Fix 𝑥 ∈ 𝐸 with 𝑥 ≠ 0 and set 𝑀 ∶= span{𝑥}. Define 𝑓0 ∈ Hom(𝑀, 𝕂) by
setting ⟨𝑓0 , 𝛼𝑥⟩ ∶= 𝛼‖𝑥‖𝐸 for all 𝛼 ∈ 𝕂. Setting 𝑝(𝑦) ∶= ‖𝑦‖ we have
for all 𝛼 ∈ 𝕂. By Theorem 26.4 there exists 𝑓 ∈ Hom(𝐸, 𝕂) such that 𝑓 (𝛼𝑥) = 𝑓0 (𝛼𝑥)
for all 𝛼 ∈ 𝕂 and |⟨𝑓 , 𝑦⟩| ≤ 𝑝(𝑦) = ‖𝑦‖𝐸 for all 𝑦 ∈ 𝐸. Hence 𝑓 ∈ 𝐸 ′ and ‖𝑓 ‖𝐸 ′ ≤ 1.
Since |⟨𝑓 , 𝑥⟩| = ‖𝑥‖𝐸 by definition of 𝑓 we have ‖𝑓 ‖𝐸 ′ = 1, completing the proof of
the corollary.
26.6 Remark By the above corollary 𝐸 ′ ≠ {0} for all normed spaces 𝐸 ≠ {0}.
Theorem 26.4 also allows us to get continuous extensions of linear functionals.
so that ‖𝑓 ‖𝐸 ′ = ‖𝑓0 ‖𝑀 ′ .
We can use the above theorem to obtain linear functionals with special properties.
26.8 Corollary Let 𝐸 be a normed space and 𝑀 ⊂ 𝐸 a proper closed subspace. Then
for every 𝑥0 ∈ 𝐸 ⧵ 𝑀 there exists 𝑓 ∈ 𝐸 ′ such that 𝑓 |𝑀 = 0 and ⟨𝑓 , 𝑥0 ⟩ = 1.
92
Proof. Fix 𝑥0 ∈ 𝐸 ⧵ 𝑀 be arbitrary and let 𝑀1 ∶= 𝑀 ⊕ span{𝑥0 }. For 𝑚 + 𝛼𝑥0 ∈ 𝑀1
define
⟨𝑓0 , 𝑚 + 𝛼𝑥0 ⟩ ∶= 𝛼.
for all 𝛼 ∈ 𝕂 and 𝑚 ∈ 𝑀. Then clearly ⟨𝑓0 , 𝑥0 ⟩ = 1 and ⟨𝑓0 , 𝑚⟩ = 0 for all 𝑚 ∈ 𝑀.
Moreover, since 𝑀 is closed 𝑑 ∶= dist(𝑥0 , 𝑀) > 0 and
Hence
|⟨𝑓0 , 𝑚 + 𝛼𝑥0 ⟩| = |𝛼| ≤ 𝑑 −1 ‖𝑚 + 𝛼𝑥0 ‖
for all 𝑚 ∈ 𝑀 and 𝛼 ∈ 𝕂. Hence 𝑓0 ∈ 𝑀1′ . Now we can apply Theorem 26.7 to
conclude the proof.
27 Reflexive Spaces
The dual of a normed space is again a normed space. Hence we can look at the dual of
that space as well. The question is whether or not the second dual coincides with the
original space.
𝐸 ′′ ∶= (𝐸 ′ )′
for all 𝑓 ∈ 𝐸 ′ . Hence we can naturally identify 𝑥 ∈ 𝐸 with ⟨⋅, 𝑥⟩ ∈ 𝐸 ′′ . With that
canonical identification we have 𝐸 ⊆ 𝐸 ′′ and
27.2 Proposition The map 𝑥 → ⟨⋅, 𝑥⟩ is an isometric embedding of 𝐸 into 𝐸 ′′ , that is,
𝐸 ⊆ 𝐸 ′′ with equal norms.
For some spaces there is equality.
93
27.4 Remark (a) As dual spaces are always complete (see Remark 25.2(a)), every re-
flexive space is a Banach space.
(b) Not every Banach space is reflexive. We have shown in Example 25.4 that
𝑐0 = 𝓁1 and 𝓁1′ = 𝓁∞ , so that
′
𝑐0′′ = 𝓁1′ = 𝓁∞ ≠ 𝑐0
Other examples of non-reflexive Banach spaces are 𝐿1 (Ω), 𝐿∞ (Ω) and 𝐵𝐶(Ω).
28 Weak convergence
So far, in a Banach space we have only looked at sequences converging with respect to
the norm in the space. Since we work in infinite dimensions there are more notions of
convergence of a sequence. We introduce one involving dual spaces.
28.1 Definition (weak convergence) Let 𝐸 be a Banach space and (𝑥𝑛 ) a sequence in
𝐸. We say (𝑥𝑛 ) converges weakly to 𝑥 in 𝐸 if
lim ⟨𝑓 , 𝑥𝑛 ⟩ = ⟨𝑓 , 𝑥⟩
𝑛→∞
28.2 Remark (a) Clearly, if 𝑥𝑛 → 𝑥 with respect to the norm, then 𝑥𝑛 ⇀ 𝑥 weakly.
(b) The converse is not true, that is, weak convergence does not imply convergence
with respect to the norm. As an example consider the sequence 𝑒𝑛 ∶= (𝛿𝑘𝑛 )𝑘∈ℕ ∈ 𝓁2 .
Then |𝑒𝑛 |2 = 1 and ⟨𝑓 , 𝑒𝑛 ⟩ = 𝑓𝑛 for all 𝑓 = (𝑓𝑘 )𝑘∈ℕ ∈ 𝓁2′ = 𝓁2 . Hence for all 𝑓 ∈ 𝓁2′
⟨𝑓 , 𝑒𝑛 ⟩ = 𝑓𝑛 → 0
94
Proof. We consider 𝑥𝑛 as an element of 𝐸 ′′ . By assumption ⟨𝑓 , 𝑥𝑛 ⟩ → ⟨𝑓 , 𝑥⟩ for all
𝑓 ∈ 𝐸 ′ . Hence by the uniform boundedness principle (Corollary 22.3) and Proposi-
tion 27.2
‖𝑥‖𝐸 = ‖𝑥‖𝐸 ′′ ≤ lim inf ‖𝑥𝑛 ‖𝐸 ′′ = lim inf ‖𝑥𝑛 ‖𝐸 < ∞.
𝑛→∞ 𝑛→∞
as claimed.
29 Dual Operators
For every linear operator 𝑇 from a normed space 𝐸 into a normed space 𝐹 we can
define an operator from the dual 𝐹 ′ into 𝐸 ′ by composing a linear functional on 𝑓 with
𝑇 to get a linear functional on 𝐸.
29.1 Definition Suppose that 𝐸 and 𝐹 are Banach spaces and that 𝑇 ∈ (𝐸, 𝐹 ). We
define the dual operator 𝑇 ′ ∶ 𝐹 ′ → 𝐸 ′ by
𝑇 ′ 𝑓 ∶= 𝑓 ◦𝑇
for all 𝑓 ∈ 𝐹 ′ .
Note that for 𝑓 ∈ 𝐹 ′ and 𝑥 ∈ 𝐸 by definition
⟨𝑇 ′ 𝑓 , 𝑥⟩ = ⟨𝑓 , 𝑇 𝑥⟩.
𝑇
𝐸 𝐹
𝑓
𝑇 ′ 𝑓 = 𝑓 ◦𝑇
𝕂
29.2 Remark We can also look at the second dual of an operator 𝑇 ∈ (𝐸, 𝐹 ). Then
𝑇 ′′ = (𝑇 ′ )′ ∶ 𝐸 ′′ → 𝐹 ′′ . By definition of 𝑇 ′′ and the canonical embedding of 𝐸 ⊆ 𝐸 ′′
we always haven 𝑇 ′′ |𝐸 = 𝑇 . If 𝐸 and 𝐹 are reflexive, then 𝑇 ′′ = 𝑇 .
We next show that an operator and its dual have the same norm.
29.3 Theorem Let 𝐸 and 𝐹 be normed spaces and 𝑇 ∈ (𝐸, 𝐹 ). Then 𝑇 ′ ∈ (𝐹 ′ , 𝐸 ′ )
and ‖𝑇 ‖(𝐸,𝐹 ) = ‖𝑇 ′ ‖(𝐹 ′ ,𝐸 ′ ) .
Proof. By definition of the dual norm we have
95
so ‖𝑇 ′ ‖(𝐹 ′ ,𝐸 ′ ) ≤ ‖𝑇 ‖(𝐸,𝐹 ) . For the opposite inequality we use the Hahn-Banach The-
orem (Corollary 26.5) which guarantees that for every 𝑥 ∈ 𝐸 there exists 𝑓 ∈ 𝐹 ′ such
that ‖𝑓 ‖𝐹 ′ = 1 and ⟨𝑓 , 𝑇 𝑥⟩ = ‖𝑇 𝑥‖𝐹 . Hence, for that choice of 𝑓 we have
‖𝑇 𝑥‖𝐹 = ⟨𝑓 , 𝑇 𝑥⟩ = ⟨𝑇 ′ 𝑓 , 𝑥⟩ ≤ ‖𝑇 ′ 𝑓 ‖𝐸 ′ ‖𝑥‖𝐸
≤ ‖𝑇 ′ ‖(𝐹 ′ ,𝐸 ′ ) ‖𝑓 ‖𝐹 ′ ‖𝑥‖𝐸 = ‖𝑇 ′ ‖(𝐹 ′ ,𝐸 ′ ) ‖𝑥‖𝐸 .
Hence ‖𝑇 ′ ‖(𝐹 ′ ,𝐸 ′ ) ≥ ‖𝑇 ‖(𝐸,𝐹 ) and the assertion of the theorem follows.
We can use the dual operator to obtain properties of the original linear operator.
29.4 Proposition Let 𝐸 and 𝐹 be normed spaces and 𝑇 ∈ (𝐸, 𝐹 ). Then the image
of 𝑇 is dense in 𝐹 if and only if ker 𝑇 ′ = {0}.
Proof. Suppose that im(𝑇 ) = 𝐹 and that 𝑓 ∈ ker 𝑇 ′ . Then
0 = ⟨𝑇 ′ 𝑓 , 𝑥⟩ = ⟨𝑓 , 𝑇 𝑥⟩
for all 𝑥 ∈ 𝐸. Since im(𝑇 ) is dense in 𝐹 we conclude that 𝑓 = 0, so ker 𝑇 ′ = {0}. We
prove the converse by contrapositive, assuming that 𝑀 ∶= im(𝑇 ) is a proper subspace
of 𝐹 . By the Hahn-Banach Theorem (Corollary 26.8) there exists 𝑓 ∈ 𝐹 ′ with ‖𝑓 ‖𝐹 ′ =
1 and ⟨𝑓 , 𝑦⟩ = 0 for all 𝑦 ∈ 𝑀. In particular
⟨𝑇 ′ 𝑓 , 𝑥⟩ = ⟨𝑓 , 𝑇 𝑥⟩ = 0
for all 𝑥 ∈ 𝐸. Hence 𝑇 ′ 𝑓 = 0 but 𝑓 ≠ 0 and so ker 𝑇 ′ ≠ {0} as claimed.
We apply the above to show that the dual of an isomorphism is also an isomorphism.
29.6 Definition (continuous embedding) Suppose that 𝐸 and 𝐹 are normed spaces
and that 𝐸 ⊆ 𝐹 . If the natural injection 𝑖 ∶ 𝐸 → 𝐹 , 𝑥 → 𝑖(𝑥) ∶= 𝑥 is continuous,
then we write 𝐸 → 𝐹 . We say 𝐸 is continuously embedded into 𝐹 . If in addition 𝐸 is
𝑑
dense in 𝐹 , then we write 𝐸 → 𝐹 and say 𝐸 is densely embedded into 𝐹 .
If 𝐸 → 𝐹 and 𝑖 ∈ (𝐸, 𝐹 ) is the natural injection, then the dual is given by
⟨𝑖′ (𝑓 ), 𝑥⟩ = ⟨𝑓 , 𝑖(𝑥)⟩
for all 𝑓 ∈ 𝐹 ′ and 𝑥 ∈ 𝐸. Hence 𝑖′ (𝑓 ) = 𝑓 |𝐸 is the restriction of 𝑓 to 𝐸. Hence 𝐹 ′
𝑑
in some sense is a subset of 𝐸 ′ . However, 𝑖′ is an injection if and only if 𝐸 → 𝐹 by
Proposition 29.4. Hence we can prove the following proposition.
96
𝑑
29.7 Proposition Suppose that 𝐸 → 𝐹 . Then 𝐹 ′ → 𝐸 ′ . If 𝐸 and 𝐹 are reflexive,
𝑑
then 𝐹 ′ → 𝐸 ′ .
Proof. The first statement follows from the comments before. If 𝐸 and 𝐹 are reflexive,
then 𝑖′′ = 𝑖 by Remark 29.2. Hence (𝑖′ )′ = 𝑖 is injective, so by Proposition 29.4 𝑖′ has
𝑑
dense range, that is, 𝐹 ′ → 𝐸 ′ .
𝐽 (𝑦) ∶= (⋅ ∣ 𝑦) ∈ 𝐻 ′
30.1 Proposition The map 𝐽 ∶ 𝐻 → 𝐻 ′ defined above has the following properties:
(iii) 𝐽 ∶ 𝐻 → 𝐻 ′ is continuous.
Proof. (i) is obvious from the definition of 𝐽 and the properties of an inner product.
(ii) follows from Corollary 15.6 and the definition of the dual norm. Finally (iii) follows
from (ii) since ‖𝐽 (𝑥) − 𝐽 (𝑦)‖𝐻 ′ = ‖𝐽 (𝑥 − 𝑦)‖𝐻 ′ = ‖𝑥 − 𝑦‖𝐻 .
𝐻 = ker 𝑓 ⊕ (ker 𝑓 )⟂
97
for all 𝑥 ∈ 𝐻. Since 𝑧 ∈ (ker 𝑓 )⟂ we get
( | )
0 = −⟨𝑓 , 𝑥⟩𝑧 + ⟨𝑓 , 𝑧⟩𝑥|𝑧 = −⟨𝑓 , 𝑥⟩(𝑧 ∣ 𝑧) + ⟨𝑓 , 𝑧⟩(𝑥 ∣ 𝑧) = −⟨𝑓 , 𝑥⟩ + ⟨𝑓 , 𝑧⟩(𝑥 ∣ 𝑧)
|
for all 𝑥 ∈ 𝐻. Hence
( | )
⟨𝑓 , 𝑥⟩ = ⟨𝑓 , 𝑧⟩(𝑥 ∣ 𝑧) = 𝑥|⟨𝑓 , 𝑧⟩𝑧
|
for all 𝑥 ∈ 𝐻, so 𝐽 (𝑦) = 𝑓 if we set 𝑦 = ⟨𝑓 , 𝑧⟩𝑧.
We use the above to show that Hilbert spaces are reflexive.
30.3 Corollary (reflexivity of Hilbert spaces) If 𝐻 is a Hilbert space, then 𝐻 ′ is a
Hilbert space with inner product (𝑓 ∣ 𝑔)𝐻 ′ ∶= (𝐽 −1 (𝑔) ∣ 𝐽 −1 (𝑓 ))𝐻 . The latter inner
product induces the the dual norm. Moreover, 𝐻 is reflexive.
Proof. By the properties of 𝐽 , (𝑓 ∣ 𝑔)𝐻 ′ ∶= (𝐽 −1 (𝑔) ∣ 𝐽 −1 (𝑓 ))𝐻 clearly defines an
inner product on 𝐻 ′ . Moreover, again by the properties of 𝐽
(𝑓 ∣ 𝑓 )𝐻 ′ = (𝐽 −1 (𝑓 ) ∣ 𝐽 −1 (𝑓 ))𝐻 = ‖𝑓 ‖2𝐻 ′ ,
so 𝐻 ′ is a Hilbert space. To prove that 𝐻 is reflexive let 𝑥 ∈ 𝐻 ′′ and denote the Riesz
isomorphisms on 𝐻 and 𝐻 ′ by 𝐽𝐻 and 𝐽𝐻 ′ , respectively. Then by the definition of the
inner product on 𝐻 ′ and 𝐽𝐻
⟨𝑓 , 𝑥⟩𝐻 = ⟨𝑥, 𝑓 ⟩𝐻 ′ = (𝑓 ∣ 𝐽𝐻−1′ (𝑥))𝐻 ′ = (𝐽𝐻−1 𝐽𝐻−1′ (𝑥) ∣ 𝐽𝐻−1 (𝑓 ))𝐻 = ⟨𝑓 , 𝐽𝐻−1 𝐽𝐻−1′ (𝑥)⟩
for all 𝑓 ∈ 𝐻 ′ . Hence by definition of reflexivity 𝐻 is reflexive.
Let finally 𝐻1 and 𝐻2 be Hilbert spaces and 𝑇 ∈ (𝐻1 , 𝐻2 ). Moreover, let 𝐽1 and 𝐽2
be the Riesz isomorphisms on 𝐻1 and 𝐻2 , respectively. If 𝑇 ′ is the dual operator of 𝑇 ,
then we set
𝑇 ∗ = 𝐽1−1 ◦𝑇 ′ ◦𝐽2 ∈ (𝐻2 , 𝐻1 ).
The situation is depicted in the commutative diagram below:
𝑇′
𝐻2′ 𝐻1′
𝐽2 𝐽1
𝑇∗
𝐻2 𝐻1
98
31 The Lax-Milgram Theorem
In the previous section we looked at the representation of linear functionals on a Hilbert
space by means of the scalar product. Now we look at sesqui-linear forms.
31.2 Proposition For every bounded sesqui-linear form 𝑏(⋅ , ⋅) there exists a unique
linear operator 𝐴 ∈ (𝐻) such that (31.3) holds for all 𝑢, 𝑣 ∈ 𝐻. Moreover, ‖𝐴‖(𝐻) ≤
𝑀, where 𝑀 is the constant in (31.1).
Proof. By (31.1) it follows that for every fixed 𝑣 ∈ 𝐻 the map 𝑢 → 𝑏(𝑢, 𝑣) is a bounded
linear functional on 𝐻. The Riesz representation theorem implies that there exists a
unique element 𝐴𝑣 in 𝐻 such that (31.3) holds. The map 𝑣 → 𝐴𝑣 is linear because
̄
(𝑤 ∣ 𝐴(𝜆𝑢 + 𝜇𝑣)) = 𝑏(𝑤, 𝜆𝑢 + 𝜇𝑣) = 𝜆𝑏(𝑤, 𝑢) + 𝜇𝑏(𝑤,
̄ 𝑣)
̄ ∣ 𝐴𝑢) + 𝜇(𝑤
= 𝜆(𝑤 ̄ ∣ 𝐴𝑣) = (𝑤 ∣ 𝜆𝐴𝑢 + 𝜇𝐴𝑣)
for all 𝑢, 𝑣, 𝑤 ∈ 𝐻 and 𝜆, 𝜇 ∈ 𝕂, and so by definition of 𝐴 we get 𝐴(𝜆𝑢 + 𝜇𝑣) =
𝜆𝐴𝑢 + 𝜇𝐴𝑣. To show that 𝐴 is bounded we note that by Corollary 15.6 and (31.1) we
have
‖𝐴𝑣‖𝐻 = sup |(𝑢 ∣ 𝐴𝑣)| = sup |𝑏(𝑢, 𝑣)| ≤ sup 𝑀‖𝑢‖𝐻 ‖𝑣‖𝐻 = 𝑀‖𝑣‖𝐻
‖𝑢‖𝐻 =1 ‖𝑢‖𝐻 =1 ‖𝑢‖𝐻 =1
We next show that if the form is coercive, then the operator constructed above is in-
vertible.
99
31.3 Theorem (Lax-Milgram theorem) Suppose 𝑏(⋅ , ⋅) is a bounded coercive sesqui-
linear form and 𝐴 ∈ (𝐻) the corresponding operator constructed above. Then 𝐴 is
invertible and 𝐴−1 ∈ (𝐻) with ‖𝐴−1 ‖(𝐻) ≤ 𝛼 −1 , where 𝛼 > 0 is the constant from
(31.2).
Proof. By the coercivity of 𝑏(⋅ , ⋅) we have
by definition of 𝐴 and the coercivity of 𝑏(⋅ , ⋅). Hence 𝑢 = 0 and so Corollary 16.11 im-
plies that im(𝐴) is dense in 𝐻. Since 𝐴 is in particular a closed operator with bounded
inverse, Theorem 23.4(iii) implies that im(𝐴) = 𝐻.
100
Chapter VII
Spectral Theory
32.1 Definition (resolvent and spectrum) Let 𝐸 be a Banach space over ℂ and 𝐴 ∶ 𝐷(𝐴) ⊆
𝐸 → 𝐸 a closed operator. We call
and
101
We next get some characterisations of the points in the resolvent set.
exists and 𝑟 ≤ ‖𝑇 ‖.
Proof. First note that
for all 𝑛, 𝑚 ≥ 0. Fix now 𝑚 ∈ ℕ, 𝑚 > 1. If 𝑛 > 𝑚, then there exist 𝑞𝑛 , 𝑟𝑛 ∈ ℕ with
𝑛 = 𝑚𝑞𝑛 + 𝑟𝑛 and 0 ≤ 𝑟𝑛 < 𝑚. Hence
𝑞𝑛 ( 𝑟 )
1 1
= 1+ 𝑛 →
𝑛 𝑚 𝑛 𝑚
as 𝑛 → ∞. From (32.2) we conclude that ‖𝑇 𝑛 ‖1∕𝑛 ≤ ‖𝑇 𝑚 ‖𝑞𝑛 ∕𝑛 ‖𝑇 ‖𝑟𝑛 ∕𝑛 for all 𝑛 > 𝑚.
Hence
lim inf ‖𝑇 𝑛 ‖1∕𝑛 ≤ ‖𝑇 𝑚 ‖1∕𝑚
𝑛→∞
We now look at an analogue of the “geometric series” for bounded linear operators.
102
32.5 Proposition (Neumann series) Let 𝑇 ∈ (𝐸) and set
𝑟 ∶= lim ‖𝑇 𝑘 ‖1∕𝑘 .
𝑘→∞
Proof. Part (i) is a consequence of the root test for the absolute convergence of a series.
For part (ii) we use a similar argument as in case of the geometric series. Clearly
∑
𝑛
∑
𝑛
∑
𝑛
∑
𝑛+1
𝑘 𝑘 𝑘
(𝐼 − 𝑇 ) 𝑇 = 𝑇 (𝐼 − 𝑇 ) = 𝑇 − 𝑇 𝑘 = 𝐼 − 𝑇 𝑛+1
𝑘=0 𝑘=0 𝑘=0 𝑘=0
∑𝑛
for all 𝑛 ∈ ℕ. Since the series 𝑘=0
𝑇 𝑘 converges, in particular 𝑇 𝑛 → 0 in (𝐸) and
therefore
∑
∞
∑
∞
(𝐼 − 𝑇 ) 𝑇 𝑘𝑥 = 𝑇 𝑘 (𝐼 − 𝑇 )𝑥 = 𝑥
𝑘=0 𝑘=0
We next use the above to show that the resolvent is analytic on 𝜚(𝐴). The series expan-
sions we find are very similar to corresponding expansions of (𝜆 − 𝑎)−1 based on the
geometric series.
∑
∞
𝑓 (𝜆) = 𝑎𝑘 (𝜆 − 𝜆0 )𝑘
𝑘=0
(i) The resolvent set 𝜚(𝐴) is open in ℂ and the resolvent 𝑅(⋅ , 𝐴) ∶ 𝜚(𝐴) → (𝐸) is
analytic;
1
(ii) For all 𝜆0 ∈ 𝜚(𝐴) we have ‖𝑅(𝜆0 , 𝐴)‖(𝐸) ≥ .
dist(𝜆0 , 𝜎(𝐴))
(iii) For all 𝜆, 𝜇 ∈ 𝜚(𝐴)
(resolvent equation);
103
(iv) For all 𝜆, 𝜇 ∈ 𝜚(𝐴)
and
𝐴𝑅(𝜆, 𝐴)𝑥 = 𝑅(𝜆, 𝐴)𝐴𝑥 = (𝜆𝑅(𝜆, 𝐴) − 𝐼)𝑥
for all 𝑥 ∈ 𝐷(𝐴).
Proof. (i) If 𝜚(𝐴) = ∅, then 𝜚(𝐴) is open. Hence suppose that 𝜚(𝐴) ≠ ∅ and fix
𝜆0 ∈ 𝜚(𝐴). Then
𝜆𝐼 − 𝐴 = (𝜆 − 𝜆0 ) + (𝜆0 𝐼 − 𝐴)
for all 𝜆 ∈ ℂ. Hence
( )
(𝜆𝐼 − 𝐴) = (𝜆0 𝐼 − 𝐴) 𝐼 + (𝜆 − 𝜆0 )𝑅(𝜆0 , 𝐴) (32.3)
Hence by (32.3) the operator (𝜆𝐼 − 𝐴) ∶ 𝐷(𝐴) → 𝐸 has a bounded inverse and
( )−1
𝑅(𝜆, 𝐴) = (𝜆𝐼 − 𝐴)−1 = 𝐼 + (𝜆 − 𝜆0 )𝑅(𝜆0 , 𝐴) 𝑅(𝜆0 , 𝐴)
∑
∞
= (−1)𝑘 𝑅(𝜆0 , 𝐴)𝑘+1 (𝜆 − 𝜆0 )𝑘
𝑘=0
104
If we multiply by 𝑅(𝜆, 𝐴) from the left and by 𝑅(𝜇, 𝐴) from the right we get
𝐴 = 𝜆𝐼 − (𝜆𝐼 − 𝐴)
32.8 Remarks (a) Because 𝜚(𝐴) is open, the spectrum 𝜎(𝐴) is always a closed subset
of ℂ.
(b) Part (ii) if the above theorem shows that 𝜚(𝐴) is the natural domain of the analytic
function given by the resolvent 𝑅(⋅ , 𝐴). The function is Banach space valued, but
nevertheless all theorems from complex analysis can be applied. In particular, one can
analyse the structure of 𝐴 near an isolated point of the spectrum by using Laurent series
expansions, see for instance [3].
Next we look at the spectrum of bounded operators.
32.10 Theorem Suppose that 𝐸 is a Banach space and 𝑇 ∈ (𝐸). Then 𝜎(𝑇 ) is non-
empty and bounded. Moreover, spr(𝑇 ) = lim𝑛→∞ ‖𝑇 𝑛 ‖1∕𝑛 ≤ ‖𝑇 ‖.
Proof. Using Proposition 32.5 we see that
( )
1 −1 ∑ 1 𝑘
∞
1
𝑅(𝜆, 𝑇 ) = (𝜆𝐼 − 𝑇 )−1 = 𝐼− 𝑇 = 𝑘+1
𝑇 (32.4)
𝜆 𝜆 𝑘=0
𝜆
if |𝜆| > 𝑟 ∶= lim𝑛→∞ ‖𝑇 𝑛 ‖1∕𝑛 . The above is the Laurent expansion of 𝑅(⋅ , 𝑇 ) about
zero for |𝜆| large. We know that such an expansion converges in the largest annulus
contained in the domain of the analytic function. Hence 𝑟 = spr(𝑇 ) as claimed. We
next prove that 𝜎(𝑇 ) is non-empty. We assume 𝜎(𝑇 ) is empty and derive a contradiction.
If 𝜎(𝑇 ) = ∅, then the function
𝑔 ∶ ℂ → ℂ, 𝜆 → ⟨𝑓 , 𝑅(𝜆, 𝑇 )𝑥⟩
1 ∑ ‖𝑇 ‖𝑘
∞
‖𝑓 ‖𝐸 ′ ‖𝑥‖𝐸
|𝑔(𝜆)| ≤ ‖𝑓 ‖ ′ ‖𝑥‖𝐸 =
|𝜆| 𝑘=0 2𝑘 ‖𝑇 ‖𝑘 𝐸
|𝜆|
105
for all |𝜆| ≥ 2‖𝑇 ‖. In particular, 𝑔 is bounded and therefore by Liouville’s theorem 𝑔
is constant. Again from the above 𝑔(𝜆) → 0 as |𝜆| → ∞, and so that constant must be
zero. Hence
⟨𝑓 , 𝑅(𝜆, 𝑇 )𝑥⟩ = 0
for all 𝑥 ∈ 𝐸 and 𝑓 ∈ 𝐸 ′ . Hence by the Hahn-Banach theorem (Corollary 26.5)
𝑅(𝜆, 𝑇 )𝑥 = 0 for all 𝑥 ∈ 𝐸, which is impossible since 𝑅(𝜆, 𝑇 ) is bijective. This proves
that 𝜎(𝑇 ) ≠ ∅.
Note that 𝜆𝐼 − 𝐴 may fail to have a continuous inverse defined on 𝐸 for several reasons.
We split the spectrum up accordingly.
106
33.2 Lemma (i) If 𝑃 ∈ Hom(𝐸) is a projection, then 𝐸 = im(𝑃 ) ⊕ ker(𝑃 ). Moreover,
𝐼 − 𝑃 is a projection and ker(𝑃 ) = im(𝐼 − 𝑃 ).
(ii) If 𝐸1 , 𝐸2 are subspaces of 𝐸 with 𝐸 = 𝐸1 ⊕ 𝐸2 , then there exists a unique
projection 𝑃 ∈ Hom(𝐸) with 𝐸1 = im(𝑃 ) and 𝐸2 = im(𝐼 − 𝑃 ).
Proof. (i) If 𝑥 ∈ 𝐸, then since 𝑃 2 = 𝑃
𝑃 (𝐼 − 𝑃 )𝑥 = 𝑃 𝑥 − 𝑃 2 𝑥 = 𝑃 𝑥 − 𝑃 𝑥 = 0,
𝐸2
𝑥
𝑃)
−
𝑥
(𝐼
𝑃𝑥
𝐸1
107
33.4 Definition (topological direct sum) Suppose that 𝐸1 , 𝐸2 are subspaces of the normed
space 𝐸. We call 𝐸 = 𝐸1 ⊕ 𝐸2 a topological direct sum if the corresponding projection
is bounded. In that case we say 𝐸2 is a topological complement of 𝐸1 . Finally, a closed
subspace of 𝐸 is called complemented if it has a topological complement.
33.6 Lemma Suppose that 𝑀 is a closed subspace of the Banach space 𝐸 with finite
co-dimension, that is, dim(𝐸∕𝑀) < ∞. Then 𝑀 is complemented.
Proof. We know that 𝑀 has an algebraic complement 𝑁, that is, 𝐸 = 𝑀 ⊕ 𝑁. Let
𝑃 denote the projection onto 𝑁. Then we have the following commutative diagram.
𝑃
𝐸 𝑁
𝜋
𝑃̂
𝐸∕𝑀
Suppose now that 𝐸 is a vector space and 𝐸1 and 𝐸2 are subspaces such that 𝐸 =
𝐸1 ⊕ 𝐸2 . Denote by 𝑃1 and 𝑃2 the corresponding projections. If 𝐴 ∈ Hom(𝐸), then
𝐴𝑥 = 𝑃1 𝐴𝑥 + 𝑃2 𝐴𝑥 = 𝑃1 𝐴(𝑃1 𝑥 + 𝑃2 𝑥) + 𝑃2 𝐴(𝑃1 𝑥 + 𝑃2 𝑥)
= 𝑃1 𝐴𝑃1 𝑥 + 𝑃1 𝐴𝑃2 𝑥 + 𝑃2 𝐴𝑃1 𝑥 + 𝑃2 𝐴𝑃2 𝑥
108
33.7 Definition (complete reduction) We say 𝐸 = 𝐸1 ⊕ 𝐸2 completely reduces 𝐴 if
𝐴12 = 𝐴21 = 0, that is, [ ]
𝐴11 0
𝐴∼
0 𝐴22
If that is the case we set 𝐴𝑖 ∶= 𝐴𝑖𝑖 and write 𝐴 = 𝐴1 ⊕ 𝐴2 .
We next characterise the reductions by means of the projections.
33.8 Proposition Suppose that 𝐸 = 𝐸1 ⊕𝐸2 and that 𝑃1 and 𝑃2 are the corresponding
projections. Let 𝐴 ∈ Hom(𝐸). Then 𝐸 = 𝐸1 ⊕ 𝐸2 completely reduces 𝐴 if and only if
𝑃1 𝐴 = 𝐴𝑃1 .
Proof. Suppose that 𝐸 = 𝐸1 ⊕ 𝐸2 completely reduces 𝐴. Then
Of course we could replace interchange the roles of 𝑃1 and 𝑃2 in the above proposition.
(ii) im 𝐴 = im 𝐴1 ⊕ im 𝐴2 ;
We finally look at resolvent and spectrum of the reductions. It turns out that the spectral
properties of 𝐴 can can be completely described by the operators 𝐴1 and 𝐴2 .
109
(iii) 𝜎𝑝 (𝐴) = 𝜎𝑝 (𝐴1 ) ∪ 𝜎𝑝 (𝐴2 ).
Proof. By Proposition 33.3 the projections 𝑃1 , 𝑃2 associated with 𝐸 = 𝐸1 ⊕ 𝐸2 are
bounded. Hence 𝐴𝑖 = 𝑃𝑖 𝐴𝑃𝑖 is bounded for 𝑖 = 1, 2. The equality in (ii) follows from
Proposition 33.9(iv) which shows that 𝜚(𝐴) = 𝜚(𝐴1 ) ∩ 𝜚(𝐴2 ). Finally (iii) follows since
by Proposition 33.9(i) shows that ker(𝜆𝐼 − 𝐴) = {0} if and only if ker(𝜆𝐼𝐸1 − 𝐴1 ) =
ker(𝜆𝐼𝐸2 − 𝐴2 ) = {0}.
and
𝐸 = im 𝑇 0 ⊇ im 𝑇 ⊇ im 𝑇 2 ⊇ im 𝑇 3 ⊇ … .
We are interested in 𝑛 such that there is equality.
34.1 Lemma If ker 𝑇 𝑛 = ker 𝑇 𝑛+1 , then ker 𝑇 𝑛 = ker 𝑇 𝑛+𝑘 for all 𝑘 ∈ ℕ. Moreover,
if im 𝑇 𝑛 = im 𝑇 𝑛+1 , then im 𝑇 𝑛 = im 𝑇 𝑛+𝑘 for all 𝑘 ∈ ℕ.
Proof. If 𝑘 ≥ 1 and 𝑥 ∈ ker 𝑇 𝑛+𝑘 , then 0 = 𝑇 𝑛+𝑘 𝑥 = 𝑇 𝑛+1 (𝑇 𝑘−1 𝑥) and so 𝑇 𝑘−1 𝑥 ∈
ker 𝑇 𝑛 . By assumption ker 𝑇 𝑛 = ker 𝑇 𝑛+1 and therefore 𝑇 𝑛+𝑘−1 𝑥 = 𝑇 𝑛 (𝑇 𝑘−1 𝑥) = 0.
Hence ker 𝑇 𝑛+𝑘 = ker 𝑇 𝑛+𝑘−1 for all 𝑘 ≥ 1 from which the first assertion follows.
The second assertion is proved similarly. If 𝑘 ≥ 1 and 𝑥 ∈ im 𝑇 𝑛+𝑘−1 , then there
exists 𝑦 ∈ 𝐸 with 𝑥 = 𝑇 𝑛+𝑘−1 𝑦 = 𝑇 𝑘−1 (𝑇 𝑛 𝑦) and so 𝑇 𝑛 𝑦 ∈ im 𝑇 𝑛 . By assumption
im 𝑇 𝑛 = im 𝑇 𝑛+1 and therefore there exists 𝑧 ∈ im 𝑇 𝑛+1 with 𝑇 𝑛+1 𝑧 = 𝑇 𝑛 𝑦. Thus
𝑇 𝑛+𝑘 𝑧 = 𝑇 𝑘−1 (𝑇 𝑛+1 𝑧) = 𝑇 𝑘−1 (𝑇 𝑛 𝑦) = 𝑇 𝑛+𝑘−1 𝑦 = 𝑥 which implies that im 𝑇 𝑛+𝑘−1 =
im 𝑇 𝑛+𝑘 for all 𝑘 ≥ 1. This completes the proof of the lemma.
the descent of 𝑇
If dim 𝐸 = 𝑁 < ∞, then ascent and descent are clearly finite and since dim(ker 𝑇 𝑛 ) +
dim(im(𝑇 𝑛 )) = 𝑁 for all 𝑛 ∈ ℕ, ascent and descent are always equal. We now show that
they are equal provided they are both finite. The arguments we use are similar to those
to prove the Jordan decomposition theorem for matrices. The additional complication
is that in infinite dimensions there is no dimension formula.
110
34.3 Theorem Suppose that 𝛼(𝑇 ) and 𝛿(𝑇 ) are both finite. Then 𝛼(𝑇 ) = 𝛿(𝑇 ). More-
over, if we set 𝑛 ∶= 𝛼(𝑇 ) = 𝛿(𝑇 ), then
𝐸 = ker 𝑇 𝑛 ⊕ im 𝑇 𝑛
and the above direct sum completely reduces 𝑇 .
Proof. Let 𝛼 ∶= 𝛼(𝑇 ) and 𝛿 ∶= 𝛿(𝑇 ). The plan is to show that
im 𝑇 𝛼 ∩ ker 𝑇 𝑘 = {0} (34.1)
and that
im 𝑇 𝑘 + ker 𝑇 𝛿 = 𝐸 (34.2)
for all 𝑘 ≥ 1. Choosing 𝑘 = 𝛿 in (34.1) and 𝑘 = 𝛼 in (34.2) we then get
im 𝑇 𝛼 ⊕ ker 𝑇 𝛿 = 𝐸. (34.3)
To show that 𝛼 = 𝛿 assume that 𝑥 ∈ ker 𝑇 𝛿+1 . By (34.3) there is a unique decompo-
sition 𝑥 = 𝑥1 + 𝑥2 with 𝑥1 ∈ im 𝑇 𝛼 and 𝑥2 ∈ ker 𝑇 𝛿 . Because ker 𝑇 𝛿 ⊆ ker 𝑇 𝛿+1 we
get 𝑥1 = 𝑥 − 𝑥2 ∈ ker 𝑇 𝛿+1 . We also have 𝑥1 ∈ im 𝑇 𝛼 , so by (34.1) we conclude that
𝑥1 = 0. Hence 𝑥 = 𝑥2 ∈ ker 𝑇 𝛿 , so that ker 𝑇 𝛿 = ker 𝑇 𝛿+1 . By definition of the ascent
we have 𝛼 ≤ 𝛿. Assuming that 𝛼 < 𝛿, then by definition of the descent im 𝑇 𝛿 ⊊ im 𝑇 𝛼 .
By (34.2) we still have im 𝑇 𝛿 + ker 𝑇 𝛿 = 𝐸, contradicting (34.3). Hence 𝛼 = 𝛿.
To prove (34.1) fix 𝑘 ≥ 1 and let 𝑥 ∈ im 𝑇 𝛼 ∩ ker 𝑇 𝑘 . Then 𝑥 = 𝑇 𝛼 𝑦 for some
𝑦 ∈ 𝐸 and 0 = 𝑇 𝑘 𝑥 = 𝑇 𝛼+𝑘 𝑦. As ker 𝑇 𝛼 = ker 𝑇 𝛼+𝑘 we also have 𝑥 = 𝑇 𝛼 𝑦 = 0. To
prove (34.2) fix 𝑘 ≥ 1 and let 𝑥 ∈ 𝐸. As im 𝑇 𝛿 = im 𝑇 𝛿+𝑘 there exists 𝑦 ∈ 𝐸 with
𝑇 𝛿 𝑥 = 𝑇 𝛿+𝑘 𝑦. Hence 𝑇 𝛿 (𝑥 − 𝑇 𝑘 𝑦) = 0, proving that 𝑥 − 𝑇 𝑘 𝑦 ∈ ker 𝑇 𝛿 . Hence we can
write 𝑥 = 𝑇 𝑘 𝑦 + (𝑥 − 𝑇 𝑘 𝑦) from which (34.2) follows.
We finally prove that (34.3) completely reduces 𝑇 . By definition of 𝑛 = 𝛼 = 𝛿
we have 𝑇 (ker 𝑇 𝑛 ) ⊆ ker 𝑇 𝑛+1 = ker 𝑇 𝑛 and 𝑇 (im 𝑇 𝑛 ) ⊆ im 𝑇 𝑛+1 = im 𝑇 𝑛 . This
completes the proof of the theorem.
We well apply the above to the operator 𝜆𝐼 − 𝐴, where 𝐴 is a bounded operator on a
Banach space and 𝜆 ∈ 𝜎𝑝 (𝐴).
34.4 Definition (algebraic multiplicity) Let 𝐸 be a Banach space and over ℂ. If 𝐴 ∈
(𝐸) and 𝜆 ∈ 𝜎𝑝 (𝐴) is an eigenvalue we call
dim ker(𝜆𝐼 − 𝐴)
the geometric multiplicity of 𝜆 and
(⋃ )
dim ker(𝜆𝐼 − 𝐴)𝑛
𝑛∈ℕ
the algebraic multiplicity of 𝜆. The ascent 𝛼(𝜆𝐼 − 𝐴) is called the Riesz index of the
eigenvalue 𝜆.
Note that the definitions here are consistent with the ones from linear algebra. As
it turns out the algebraic multiplicity of the eigenvalue of a matrix as defined above
coincides with the multiplicity of 𝜆 as a root of the characteristic polynomial. In infinite
dimensions there is no characteristic polynomial, so we use the above as the algebraic
multiplicity. The only question now is whether there are classes of operators for which
ascent and descent of 𝜆𝐼 − 𝐴 are finite for every or some eigenvalues. It turns out that
such properties are connected to compactness. This is the topic of the next section.
111
35 The Spectrum of Compact Operators
We want to study a class of operators appearing frequently in the theory of partial
differential equations and elsewhere. They share many properties with operators on
finite dimensional spaces.
35.1 Definition (compact operator) Let 𝐸 and 𝐹 be Banach spaces. Then 𝑇 ∈ Hom(𝐸, 𝐹 )
is called compact if 𝑇 (𝐵) is compact for all bounded sets 𝐵 ⊂ 𝐸. We set
35.2 Remark By the linearity 𝑇 is clearly compact if and only if 𝑇 (𝐵(0, 1)) is compact.
⋃
𝑚
𝑇𝑛 (𝐵) ⊆ 𝐵(𝑦𝑘 , 𝜀∕4). (35.2)
𝑘=1
⋃
𝑚
𝑇 (𝐵) ⊆ 𝐵(𝑦𝑘 , 𝜀∕2)
𝑘=1
112
and therefore
⋃
𝑚
⋃
𝑚
𝑇 (𝐵) ⊆ 𝐵(𝑦𝑘 , 𝜀∕2) ⊆ 𝐵(𝑦𝑘 , 𝜀).
𝑘=1 𝑘=1
Since 𝜀 > 0 was arbitrary 𝑇 (𝐵) is totally bounded. Since 𝐹 is complete Theorem 4.4
implies that 𝑇 (𝐵) is compact.
Since 𝑇 is compact 𝑆 is compact as well. Note that 𝑆 is the unit sphere in ker(𝐼 − 𝑇 ),
so by Theorem 11.3 the kernel of 𝐼 − 𝑇 is finite dimensional.
(ii) Set 𝑆 ∶= 𝐼 − 𝑇 . As ker 𝑆 is closed 𝑀 ∶= 𝐸∕ ker 𝑆 is a Banach space by
Theorem 12.3 with norm ‖𝑥‖𝑀 ∶= inf 𝑦∈ker 𝑆 ‖𝑦 − 𝑥‖. Now set 𝐹 ∶= im(𝑆). Let
113
𝑆̂ ∈ (𝑀, 𝐹 ) be the map induced by 𝑆. We show that 𝑆̂ −1 is bounded on im(𝑆). If
not, then there exist 𝑦𝑛 ∈ im(𝑆) such that 𝑦𝑛 → 0 but ‖𝑆̂ −1 𝑦𝑛 ‖𝑀 = 1 for all 𝑛 ∈ ℕ. By
definition of the quotient norm there exist 𝑥𝑛 ∈ 𝐸 with 1 ≤ ‖𝑥𝑛 ‖𝐸 ≤ 2 such that
̂ 𝑛 ] = 𝑥𝑛 − 𝑇 𝑥𝑛 = 𝑦𝑛 → 0.
𝑆[𝑥
Since 𝑇 is compact there exists a subsequence such that 𝑇 𝑥𝑛𝑘 → 𝑧. Hence from the
above
𝑥𝑛𝑘 = 𝑦𝑛𝑘 + 𝑇 𝑥𝑛𝑘 → 𝑧.
By the continuity of 𝑇 we have 𝑇 𝑥𝑛𝑘 → 𝑇 𝑧, so that 𝑧 = 𝑇 𝑧. Hence 𝑧 ∈ ker(𝐼 − 𝑇 ) and
[𝑥𝑛𝑘 ] → [𝑧] = [0] in 𝑀. Since ‖[𝑥𝑛 ‖𝑀 = 1 for all 𝑛 ∈ ℕ this is not possible. Hence
𝑆̂ −1 ∈ (im(𝑆), 𝑀) and by Proposition 23.4 it follows that 𝐹 = im(𝑆)̂ = im(𝐼 − 𝑇 ) is
closed.
We next prove that the ascent and descent of 𝜆𝐼 − 𝑇 is finite if 𝑇 is compact and 𝜆 ≠ 0.
35.7 Theorem Let 𝑇 ∈ (𝐸) and 𝜆 ∈ 𝕂⧵{0}. Then 𝑛 ∶= 𝛼(𝜆𝐼 −𝑇 ) = 𝛿(𝜆𝐼 −𝑇 ) < ∞
and
𝐸 = ker(𝜆𝐼 − 𝑇 )𝑛 ⊕ im(𝜆𝐼 − 𝑇 )𝑛
is a topological direct sum completely reducing 𝜇𝐼 − 𝑇 for all 𝜇 ∈ ℂ.
Proof. Replacing 𝑇 by 𝜆𝑇 we can assume without loss of generality that 𝜆 = 1.
Suppose that 𝛼(𝜆𝐼 − 𝑇 ) = ∞. Then 𝑁𝑘 ∶= ker(𝐼 − 𝑇 )𝑘 is closed in 𝐸 for every 𝑘 ∈ ℕ
and
𝑁0 ⊂ 𝑁1 ⊂ 𝑁2 ⊂ 𝑁3 ⊂ … ,
where all inclusions are proper inclusions. By Corollary 11.2 there exist 𝑥𝑘 ∈ 𝑁𝑘 such
that
1
‖𝑥𝑘 ‖ = 1 and dist(𝑥𝑘 , 𝑁𝑘−1 ) ≥
2
for all 𝑘 ≥ 1. Now
𝑇 𝑥𝑘 − 𝑇 𝑥𝓁 = (𝐼 − 𝑇 )𝑥𝓁 − (𝐼 − 𝑇 )𝑥𝑘 + 𝑥𝑘 − 𝑥𝓁
( )
= 𝑥𝑘 − 𝑥𝓁 − (𝐼 − 𝑇 )𝑥𝓁 + (𝐼 − 𝑇 )𝑥𝑘 = 𝑥𝑘 − 𝑧,
𝑀0 ⊃ 𝑀1 ⊃ 𝑀2 ⊃ 𝑀3 ⊃ … ,
114
where all inclusions are proper inclusions. By Corollary 11.2 there exist 𝑥𝑘 ∈ 𝑀𝑘 such
that
1
‖𝑥𝑘 ‖ = 1 and dist(𝑥𝑘 , 𝑀𝑘+1 ) ≥
2
for all 𝑘 ≥ 1. Now
𝑇 𝑥𝑘 − 𝑇 𝑥𝓁 = (𝐼 − 𝑇 )𝑥𝓁 − (𝐼 − 𝑇 )𝑥𝑘 − 𝑥𝑘 + 𝑥𝓁
( )
= 𝑥𝓁 − 𝑥𝑘 − (𝐼 − 𝑇 )𝑥𝓁 + (𝐼 − 𝑇 )𝑥𝑘 = 𝑥𝑘 − 𝑧,
As a consequence of the above theorem we prove that compact operators behave quite
similarly to operators on finite dimensional spaces. The following corollary would be
proved by the dimension formula in finite dimensions.
𝐸 = ker(𝜆𝐼 − 𝑇 )𝑛 ⊕ im(𝜆𝐼 − 𝑇 )𝑛
115
is a topological direct sum completely reducing 𝜇𝐼 − 𝑇 for all 𝜇 ∈ ℂ. We set 𝑁 ∶=
ker(𝜆𝐼 − 𝑇 )𝑛 and 𝑀 ∶= im(𝜆𝐼 − 𝑇 )𝑛 . Let 𝑇 = 𝑇𝑁 ⊕ 𝑇𝑀 be the reduction associated
with the above direct sum decomposition. By construction 𝜆𝐼 − 𝑇𝑁 is nilpotent, so by
Theorem 32.10 spr(𝜆𝐼−𝑇𝑁 ) = 0. In particular this means that 𝜎(𝑇𝑁 ) = {𝜆}. Moreover,
by construction, 𝜆𝐼 − 𝑇𝑀 is injective and so by Proposition 32.3 and Corollary 35.8
𝜆 ∈ 𝜚(𝑇𝑀 ). As 𝜚(𝑇𝑀 ) is open by Theorem 32.7 there exists 𝑟 > 0 such that 𝐵(𝜆, 𝑟) ⊂
𝜚(𝑇𝑀 ). By Theorem 33.10 𝜚(𝑇 ) = 𝜚(𝑇𝑁 ) ∩ 𝜚(𝑇𝑀 ), so 𝐵(𝜆, 𝑟) ⧵ {𝜆} ⊂ 𝜚(𝑇 ), showing
that 𝜆 is an isolated eigenvalue. Since the above works for every 𝜆 ∈ 𝜎(𝑇 ) ⧵ {0} the
assertion of the theorem follows.
35.10 Remarks (a) If 𝑇 ∈ (𝐸) and dim 𝐸 = ∞, then 0 ∈ 𝜎(𝑇 ). If not, then by
Proposition 35.4 the identity operator 𝐼 = 𝑇 𝑇 −1 is compact. Hence the unit sphere in
𝐸 is compact and so by Theorem 11.3 𝐸 is finite dimensional.
(b) The above theorem allows us to decompose a compact operator 𝑇 ∈ (𝐸) in
the following way. Let 𝜆𝑘 ∈ 𝜎𝑝 (𝑇 ) ⧵ {0}, 𝑘 = 1, … , 𝑚, and let 𝑛𝑘 be the corresponding
Riesz indices. Then set
𝑁𝑘 ∶= ker(𝜆𝐼 − 𝑇 )𝑛𝑘
and 𝑇𝑘 ∶= 𝑇 |𝑁𝑘 . Then there exists a closed subspace 𝑀 ⊂ 𝐸 which is invariant under
𝑇 such that
𝐸 = 𝑁1 ⊕ 𝑁 2 ⊕ ⋯ ⊕ 𝑁 𝑚 ⊕ 𝑀
completely reduces 𝑇 = 𝑇1 ⊕ 𝑇2 ⊕ ⋯ ⊕ 𝑇𝑚 ⊕ 𝑇𝑀 . To get this reduction apply Theo-
rem 35.9 inductively.
The above theory is also useful for certain classes of closed operators 𝐴 ∶ 𝐷(𝐴) ⊆
𝐸 → 𝐸, namely those for which 𝑅(𝜆, 𝐴) is compact for some 𝜆 ∈ 𝜚(𝐴). In that case
the resolvent identity from Theorem 32.7 and Proposition 35.4 imply that 𝑅(𝜆, 𝐴) is
compact for all 𝜆 ∈ 𝜚(𝐴). It turns out that 𝜎(𝐴) = 𝜎𝑝 (𝐴) consist of isolated eigenvalues,
and the only possible point of accumulation is infinity. Examples of such a situation
include boundary value problems for partial differential equations such as the one in
Example 24.5.
116
Bibliography
[1] G. Birkhoff and G.-C. Rota. “On the completeness of Sturm-Liouville expan-
sions”. In: Amer. Math. Monthly 67 (1960), pp. 835–841. DOI: 10.2307/2309440.
[2] E. Bishop and D. Bridges. Constructive analysis. Vol. 279. Grundlehren der
Mathematischen Wissenschaften. Berlin: Springer-Verlag, 1985, pp. xii+477.
[3] A. P. Campbell and D. Daners. “Linear Algebra via Complex Analysis”. In:
Amer. Math. Monthly 120.10 (2013), pp. 877–892. DOI: 10.4169/amer.math.
monthly.120.10.877.
[4] P. Cohen. “The independence of the continuum hypothesis”. In: Proc. Nat. Acad.
Sci. U.S.A. 50 (1963), pp. 1143–1148.
[5] P. J. Cohen. “The independence of the continuum hypothesis. II”. In: Proc. Nat.
Acad. Sci. U.S.A. 51 (1964), pp. 105–110.
[6] D. Daners. “Uniform continuity of continuous functions on compact metric spaces”.
In: Amer. Math. Monthly 122.6 (2015), p. 592. DOI: 10 . 4169 / amer . math .
monthly.122.6.592.
[7] J. Dugundji. Topology. Boston, Mass.: Allyn and Bacon Inc., 1966, pp. xvi+447.
[8] K. Gödel. “The consistency of the axiom of choice and of the generalized continuum-
hypothesis”. In: Proc. Nat. Acad. Sci. U.S.A. 24 (1938), pp. 556–557.
[9] J. Lindenstrauss and L. Tzafriri. “On the complemented subspaces problem”. In:
Israel J. Math. 9 (1971), pp. 263–269.
[10] W. Rudin. Real and Complex Analysis. 2nd. New York: McGraw-Hill Inc., 1974.
[11] A. D. Sokal. “A Really Simple Elementary Proof of the Uniform Boundedness
Theorem”. In: Amer. Math. Monthly 118.5 (2011), pp. 450–452. DOI: 10.4169/
amer.math.monthly.118.05.450.
[12] M. H. Stone. “The generalized Weierstrass approximation theorem”. In: Math.
Mag. 21 (1948), pp. 167–184, 237–254. DOI: 10.2307/3029750.
[13] K. Yosida. Functional analysis. 6th. Vol. 123. Grundlehren der Mathematischen
Wissenschaften. Berlin: Springer-Verlag, 1980, pp. xii+501.
[14] E. Zermelo. “Beweis, daß jede Menge wohlgeordnet werden kann”. In: Math.
Ann. 59 (1904), pp. 514–516.
117