Anda di halaman 1dari 39

Lecture Notes on Banach Algebras

and Operator Theory

By
Dinesh J. Karia
Dedicated to:
All who critically see this Lecture Notes
Contents

Chapter 1. Introduction 1
1. A Brief History 2
2. Preliminaries 3
3. Examples 7
4. Group Algebras 14
5. New From Old 20
6. Homomorphisms 23
7. Adjoining Identity 24
8. Regular Elements 26
9. Topological Divisors of Zero 29
10. Spectrum 31
11. Gel’fand Mazur Theorems 33
12. Algebras Without Identity 34
Bibliography 36

iii
CHAPTER 1

Introduction
This course is designed in two folds. One is Banach Algebra and
the other is Operator Theory. The course will evolve along the following
guidelines listed in the syllabus of M-506.
Banach Algebras
Definition and examples of a Banach algebra – Adjoining identity
– Regular and singular elements – Topological divisors of zero – The
spectrum and spectral radius – Gel’fand-Mazur theorem – Complex
homomorphisms – Weak Topology.
Commutative Banach algebras – Maximal ideals – The radical –
Semisimplicity – Maximal ideal space– Maximal ideal space of C(X)
and A(D) (Functions continuous on the closed unit disc D and analytic
in the interior) – Wermer’s maximality theorem.
The Weierstrass approximation theorem – The Stone-Weierstrass
theorem – Involution in Banach algebra – The Gel’fand Naimark the-
orem for a commutative B ∗ -algebra.
Semisimple commutative Banach algebras – Homomorphisms of
commutative Banach algebra.
Closed ideals of C(X) – The Banach-Stone theorem – Stone-Čech
compactification.
Operator Theory
Normal, unitary, selfadjoint and positive operators – Orthogonal
projections – Invariant and reducing subspaces.
Spectra of operators – Properties of spectrum – Various parts of
spectrum – Spectral mapping theorem – Numerical range
Spectral analysis of selfadjoint and normal operators – Spectral
theorem of compact normal operators – Spectral theorem for selfad-
joint operators – Fuglede-Putnam theorem – Square root of a positive
operator – Polar decomposition of an operator
Matrix representation of an operator – Operator matrices – Direct
sum

1
2 1. INTRODUCTION

1. A Brief History
Functional Analysis has been extensively studied in the last century.
The theory aims at embarking the algebraic problems using the
Analysis. However, the work of Gel’fand in 1941 is considered
to be pioneering as far as the theory of Banach algebras is
concerned. Bonsall & Duncan [BD, p. 4] very much truly
state that “The name ‘Banach algebra’ is appropriate
only because, as a normed linear space, a Banach al-
gebra is a Banach space. However this name is too
firmly established in the literature to be changed
at this time. Given a free choice we should like
to call complete normed algebras ‘Gelfand alge-
bras’ in recognition of the distinguished pioneer-
ing work of I. M. Gel’fand in this field.” The His-
torical Notes in [Ru2, p. 373-374] makes it more
clear as to how and when the theory got the mo-
mentum. He points out that Banach’s book [Ban]
deals with the vector space structure only; he never
multiplied his objects. It was Wiener who, in 1932, dealt
with the multiplicative structure and the submultiplicativ-
ity of the norm on the Banach space of absolutely convergent
Fourier series. Stone used explicitly the ring structure of spaces
of continuous functions in his work in 1937 when he generalized the
Weierstrass approximation theorem, now known as Stone-Weierstrass
theorem. In fact, Nagumo (Japanese) first initiated the abstract study
of normed rings. But what really got this subject off the ground was
Gel’fand’s discovery (1941) of important role played the maximal ideals
of commutative algebra in his construction what is now known as the
Gel’fand transform.
The classic book [Ri] by Charles E. Rickart is like “Bible of Ba-
nach algebras.” The references [BD], [Br] [La], [Ru2] and [S] are also
worth going through. We shall mainly follow the books [S] by Simmons
and [La] by Larsen.
It was von Neumann who axiomatized Hilbert space in 1930 when
his paper laid the foundation stone for Operator Theory. Incidently,
he is the person who first thought and gave the idea of storing the
programs on computer while he was associated with the project of
2. PRELIMINARIES 3

first computer Harvard Marc. Besides, he also authored a series of


papers initiating the study of operator algebras, now well known as
von Neumann algebras. In fact, the bounded linear operators were first
explored by Hilbert himself (1910c.) The book [St] by Stone is among
the fundamental books on Operator Theory.

2. Preliminaries
In this section we develop the basic definitions and results. We
also fix up notations and conventions here for the rest of our notes.
Notations: R and C will denote respectively the fields of real numbers
and complex numbers. In our discussion K will stand for either of them.
Let us first recall the algebraic structures you have studied so far.
2.1. Definition. Let G be a nonempty set with an associative binary
operation · with the following properties.
(1) There is e ∈ G such that xe = ex = x for all x ∈ G. (e turns out
to be unique and is called identity of G.)
(2) For each x ∈ G, there is y ∈ G such that xy = yx = e. (Given an
x such a y is unique and is called the inverse of x. It is customary
to denote inverse of x by x−1 .)
Then G is called a group. Further, if for all x, y ∈ G, xy = yx, then G
is said to be commutative or abelian.
If we add one more binary operation (with some properties) in a
group, then we get a ring.
2.2. Definition. Let (R, +) be an abelian group. Further, let · be
an associative binary operation on R such that x(y + z) = xy + xz
and (y + z)x = yx + zx for all x, y, z ∈ R. Then (R, +, ·) is called a
ring. These operations are called addition and multiplication. It is a
common practice to omit the dot for the ring-multiplication. That is,
xy stands for x · y.
(1) If, in a ring R, xy = yx for all x, y ∈ R, then R is called a
commutative ring.
(2) If there is an element e in a ring R satisfying xe = ex = x for
all x ∈ R, then such an e turns out to be unique and is called
the identity or unity of R. In such a case, R is called a ring
with identity or ring with unity or unital ring. Many a times the
identity of R is denoted by 1.
4 1. INTRODUCTION

(3) Let R be a ring with unity. An element x ∈ R is said to be


invertible if there exists y ∈ R such that xy = yx = 1. Again
extending the usual notations, inverse of x is denoted by x−1 . R
is said to be a division algebra, if each nonzero element of R is
invertible.
2.3. Remark.
(1) Let R be a ring with unity and x ∈ R. Then inverse of x with
respect to + always exists. It is called the additive inverse of x
and is denoted by −x.
(2) In general, a nonzero element of a ring need not be invertible.
Now we add another (external) binary operation on a group in-
stead of the ring multiplication. This structure is called a vector space.
2.4. Definition. Let (V, +) be a commutative group. Let K denote
either R or C. Let there be an external binary operation, called scalar
multiplication, on V by K given (α, v) ∈ K × V 7→ αv ∈ V with the
following properties.
(1) (α + β)v = αv + βv for all α, β ∈ K and v ∈ V .
(2) α(u + v) = αu + αv for all α ∈ K, u, v ∈ V .
(3) (αβ)v = α(βv) for all α, β ∈ K and v ∈ V .
(4) 1v = v for all v ∈ V .
Then V is called a vector space or linear space over K. The elements
of V are called vectors and elements of K scalars.
In both of the above structures, we have combined two operations
on one set to form a structure. In such cases, the combination gives
some good results only when there is some relation between two op-
erations. For example, in a ring R, we have (y + z)x = yx + zx for
all x, y, z ∈ R. In a vector space V over K, the scalars behaves nicely
with the operation of vectors. That is, α(u + v) = αu + αv for all
α ∈ K, u, v ∈ V . Also, the vectors respect the operations of scalars;
e.g., (α + β)v = αv + βv for all α, β ∈ K and v ∈ V and (αβ)v = α(βv)
for all α, β ∈ K and v ∈ V . Now we add yet another structure called
norm on a vector space. Thus we shall have three structures on one
set.
2.5. Definition. Let X be a vector space X over K together with a
norm function k · k : X → R satisfying the following. For all x, y, z ∈ X
and α ∈ K,
2. PRELIMINARIES 5

(1) kxk ≥ 0 and kxk = 0 if and only if x = 0.


(2) kx + yk ≤ kxk + kyk.
(3) kαxk = |α|kxk.
In this case, we say that (X, k · k) is a normed linear space over K.
2.6. Remark.
(1) As all of you know from the previous course on Functional Anal-
ysis, the norm induces a metric (called the norm metric) on the
normed linear space X. If this metric is complete, then (X, k · k)
is called a Banach space.
(2) Note that a binary operation is nothing but a function. More
explicitly, addition on a normed linear space X is a mapping
(x, y) ∈ X × X 7→ x + y ∈ X; scalar multiplication on X is a
mapping (k, x) ∈ K × X 7→ kx ∈ X.
Since a normed linear space has a metric, we can talk of continuity
of addition and scalar multiplication regarded as the functions. For
this, we consider the product topologies on X × X and K × X given by
suitable metrics.
2.7. Exercise.
(1) Show that addition and scalar multiplication on a normed linear
space are continuous.
(2) Can we define a normed group? a normed ring? If yes, then
construct the definitions of the same. This is typed today.
I am honouring the Once More from the audience. Please pay
extra.
Continuing our discussion of combination of more than one struc-
tures on a set, we observe the following for a normed linear space X.
(1) A normed linear space combines two algebraic structures and one
topological structure.
(2) The metric and the operations of X are related in the following
manner.
(a) d(x, y) = d(x + z, y + z) for all x, y, z ∈ X.
(b) d(kx, ky) = |k|d(x, y) for all x, y ∈ X, k ∈ K.
To define an algebra, we couple the following operation on a single
set A.
(1) commutative group operation +;
6 1. INTRODUCTION

(2) scalar multiplication; and


(3) ring multiplication.
As pointed out earlier, to build a good theory we ought to have
some relations among all these operations. For a mathematician his/her
interest in objects is poetic; at the same time (s)he has to see that these
structures occur in nature. So, besides playing with the structures, we
must remember that structures are worth investigating only when we
have a room to use them in other sciences.

2.8. Definition. Let A be a vector space over K. Suppose there is an


associative multiplication defined on A satisfying the following.
(1) x(y + z) = xy + xz and (y + z)x = yx + zx for all x, y, z ∈ A.
(2) α(xy) = (αx)y = x(αy) for all x, y ∈ A and α ∈ K.
Then A is called an algebra over K. If K = C, then A is called a
complex algebra. If K = R, then A is called a real algebra.
(1) If, xy = yx for all x, y ∈ A, then A is called a commutative
algebra.
(2) If there is an element e in A satisfying xe = ex = x for all x ∈ A,
it is called the identity or unity of A. In such a case, A is called an
algebra with identity or an algebra with unity or a unital algebra.
(3) Let A be a unital algebra. An element x ∈ A is said to be left
invertible (respectively right invertible) if there exists y ∈ A such
that yx = 1 (respectively, xy = 1). In this case, y is called
left (respectively, right) inverse of x. An element which is left
and right invertible is called invertible. Many authors also use
the word regular instead of invertible. Again extending the usual
notations, the inverse of x is denoted by x−1 . The set of all
regular elements is denoted by G. A non-invertible element is
called singular. The set of all singular elements will be denoted
by S.

2.9. Remark. Let A be an algebra with identity and x ∈ A. The


element x can have more than one left (respectively, right) inverses.
However, if x has a left inverse y and a right inverse z, then y = z
turns out to be its unique inverse. (Prove!)

We shall not mention every time that A is an algebra over K. We


simply say that A is an algebra. In most of the cases our algebras will
3. EXAMPLES 7

be complex algebras. Finally, we add one more structure of norm on


an algebra.

2.10. Definition. Let A be an algebra. Let k · k be a norm on A


making it a normed linear space. If
kxyk ≤ kxkkyk for all x, y ∈ A, (2.10.1)
then (A, k · k) is called a normed algebra. Further, if (A, k · k) is also
a Banach space (i.e., if A is complete in the norm metric), then A is
called a Banach algebra.

Let us note that for any normed linear space X, if we define xy = 0


for all x, y ∈ X, then X becomes a normed algebra. Also X = {0}
is always a Banach algebra. We make a standing assumption that our
algebras are other than these.
Before we proceed further we note the following.

2.11. Proposition. Let A be a normed algebra. Then we have the


following.
(1) Multiplication is jointly continuous on A.
(2) If A has the identity e, then kek ≥ 1.

Proof. (1) Let {xn }, {yn } be sequences in A such that xn → x and


yn → y. Since every convergent sequence is bounded, there exists
M > 0 such that kyn k ≤ M for all n ∈ N. As a result,
kxn yn − xyk = kxn yn − xyn + xyn − xyk
≤ kxn yn − xyn k + kxyn − xyk
≤ kxn − xkkyn k + kxkkyn − yk
≤ kxn − xkM + kxkkyn − yk → 0,
giving xn yn → xy.
(2) Since our algebras do contain some non-zero elements also, we fix
kekkxk kexk kxk
x 6= 0 in A. But then kek = ≥ = = 1. 
kxk kxk kxk

3. Examples
First we note that (R, k · k) with the usual operations is a Banach
algebra over R and (C, k · k) with the usual operations is a Banach
algebra over C.
8 1. INTRODUCTION

3.1. Example. Let n ∈ N. Then A = Kn , with coordinatewise multi-


plication, is an algebra over K. Fix p ∈ [1, ∞). For x = (xi ) ∈ A, define
  p1
n
p
kxkp = Σ |xi | . Also, define kxk∞ = sup |xi |. Then (A, k · kp ),
i=1 1≤i≤n
(1 ≤ p ≤ ∞), is a Banach algebra.

Proof. We verify only the submultiplicativity of the norm on A. Fix


p ∈ [1, ∞). Let x = (xi ), y = (yi ) ∈ A. Assume that |xr | = sup |xi |.
1≤i≤n
Then
  p1   p1   p1
n n n
kxykp = Σ |xi yi |p ≤ Σ |xr yi |p = |xr | Σ |yi |p
i=1 i=1 i=1
  p1   p1
n n
p p
≤ Σ |xi | Σ |yi | = kxkp kykp (3.1.1)
i=1 i=1

In case of p = ∞, let us assume that |xt yt | = sup |xi yi | for some


1≤i≤n
t = 1, 2, . . . , n. Then
kxyk∞ = sup |xi yi | = |xt yt | = |xt ||yt | ≤ kxk∞ kyk∞ (3.1.2)
1≤i≤n

3.2. Example. For 1 ≤ p < ∞, define


∞ 1
`p = {x = (xi )∞ p p
i=1 : kxkp = ( Σ |xi | ) < ∞}. (3.2.1)
i=1

Further let
`∞ = {x = (xi )∞
i=1 : kxk∞ = sup |xi | < ∞}. (3.2.2)
i=1,2,...

For x = (xi ), y = (yi ) ∈ `p define xy = (xi yi ). Then (`p , k · kp ), (1 ≤


p ≤ ∞) are Banach algebras.

Proof. You already know that (`p , k · kp ) are Banach spaces for all
p ∈ [1, ∞]. So, we verify only the submultiplicativity of the norm
on A. Fix p ∈ [1, ∞). Let x = (xi ), y = (yi ) ∈ A. Assume first that
1 ≤ p < ∞. Clearly, the sequences are bounded as they form summable
series. Also,
  p1   p1
∞ ∞
p p
|xi | ≤ Σ |xi | for all i ⇒ kxk∞ ≤ Σ |xi | = kxkp .
i=1 i=1
3. EXAMPLES 9

Hence,
  p1

p
kxykp = Σ |xi yi |
i=1
  p1

p
≤ Σ |kxk∞ yi |
i=1
  p1

p
= kxk∞ Σ |yi |
i=1

≤ kxkp kykp . (3.2.3)


In case of p = ∞, we have
kxyk∞ = sup |xi yi | = sup |xi | sup |yi | = kxk∞ kyk∞ (3.2.4)
1≤i<∞ 1≤i<∞ 1≤i<∞

Note that the submultiplicativity of k · kp in (3.2.3) and (3.2.4) also


implies that `p is closed under the pointwise product. 
3.3. Example. Let c be the set of all convergent sequences in K. Then
(c, k · k∞ ) is a Banach algebra with pointwise multiplication.
Solution. Clearly, (pointwise) product of two convergent sequences
is again a convergent sequence. Thus c is closed under multiplication.
Further, every convergent sequence is bounded. So, c ⊂ `∞ . Thus
k · k∞ retains its submultiplicativity on the smaller set c also. Since
you know the completeness of (c, k · k∞ ), we assert that it is a Banach
algebra. 
3.4. Example. Define c0 to be the set of all sequences in K converging
to zero. Then (c0 , k · k∞ ) is a Banach algebra with pointwise operations.
Solution. For two sequences xn → 0 and yn → 0, we have xn yn → 0.
So, c0 is closed under multiplication. Also, k · k∞ retains its submulti-
plicativity on the smaller set c0 also. (prove the completeness.) 
3.5. Example. Let c00 denote the set of all sequences x = (xn ) in K
such that x has only finitely many nonzero terms. Then c00 is a normed
algebra with pointwise multiplication and k · k∞ .
Solution. Left to the reader. 
3.6. Example. For a nonempty set X, let B(X) denote the set of
all K-valued bounded functions on X. For f ∈ B(X), define kf k∞ =
sup |f (x)|. Then (B(X), k · k∞ ) is a Banach algebra.
x∈X
10 1. INTRODUCTION

Proof. We check only that B(X) is closed under multiplication, k · k∞


is submultiplicative and (B(X), k · k∞ ) is complete. Clearly, f : X →
K is bounded if and only if kf k∞ < ∞. Now for f, g ∈ B(X),

kf gk∞ = sup |f (x)g(x)| ≤ sup |f (x)| sup |g(x)| = kf k∞ kgk∞ .


x∈X x∈X x∈X

Since f, g ∈ B(X), kf k∞ kgk∞ < ∞. As a result, from the above


inequality kf gk∞ < ∞. Thus the submultiplicativity of the norm
implies that B(X) is closed under multiplication. Now let {fn } be a
Cauchy sequence in B(X). For a fixed ε > 0, choose n0 ∈ N such that
ε
kfn − fm k∞ < for all n, m > n0 (3.6.1)
2
Consequently, for all n, m ≥ n0 and for every x ∈ X,
ε
|fn (x) − fm (x)| ≤ kfn − fm k∞ < . (3.6.2)
2
Hence {fn (x)} is Cauchy (and so convergent) in K for all x ∈ X.
Define f (x) = lim fn (x), (x ∈ X). Now fixing any n ≥ n0 and taking
n→∞
m → ∞ in (3.6.2), we get,
ε
|fn (x) − f (x)| ≤ for all n ≥ n0 and for every x ∈ X. (3.6.3)
2
Taking the supremum over x ∈ X in (3.6.3), we have
ε
kfn − f k∞ ≤ < ε for all n ≥ n0 . (3.6.4)
2
(3.6.4) shows that fn − f ∈ B(X). So, f = fn − (fn − f ) ∈ B(X) and
fn → f in B(X). This completes the proof. 

3.7. Example. Let X be a topological space and Cb (X) denote all


K-valued bounded continuous functions on X. For f ∈ Cb (X) define
kf k∞ = sup |f (x)|. Then (Cb (X), k · k∞ ) is a Banach algebra.
x∈X

Proof. Since multiplication of two K-valued continuous functions is


continuous (prove!), Cb (X) is an algebra. Since Cb (X) ⊂ B(X), the
norm retains its submultiplicativity on the smaller set. Thus (Cb (X), k · k∞ )
is a normed algebra. Let f ∈ B(X) be a limit point of Cb (X) and ε > 0
be given. Choose g ∈ Cb (X) such that
ε
kg − f k∞ ≤ .
3
3. EXAMPLES 11

To show the continuity of f at x0 ∈ X, choose an open set U in X,


such that |g(x) − g(x0 )| < 3ε . But then
|f (x) − f (x0 )| ≤ |f (x) − g(x)| + |g(x) − g(x0 )| + |g(x0 ) − f (x0 )|
< 2kf − gk∞ + |g(x) − g(x0 )| < ε.

3.8. Example. Let X be a compact Hausdorff topological space. Then
C(X), the set of all K-valued continuous functions of X is a Banach
algebra with k · k∞ .
Proof. Indeed, since X is compact, each continuous K-valued func-
tion on X is bounded. That is, C(X) ⊂ Cb (X). But Cb (X) ⊂ C(X)
holds for any topological space X. Hence when X is compact,
C(X) = Cb (X). Thus (C(X), k · k∞ ) is a Banach algebra. 
3.9. Remark. In general a continuous function need not be bounded
(f : R → R; f (x) = x is not bounded). As a result, when X is not
compact, we define yet another notion to build a Banach algebra.
3.10. Definition. Let X be a topological space. A function f : X → K
is said to vanish at infinity if for each ε > 0, there exists a compact
subset E of X such that |f (x)| < ε for all x ∈
/ E.
3.11. Example. Let X be a locally compact T2 -space. Define
C0 (X) = {f ∈ C(X) : f vanishes at infinity}.
Then (C0 (X), k · k∞ ) is a Banach algebra.
Proof. Clearly, C0 (X) ⊂ Cb (X). So we need only to check that
C0 (X) is closed under multiplication. Let f, g ∈ C0 (X). Let ε > 0.
Let E and F be two compact subsets of X such that |f (x)| < ε for all
x∈/ E and |g(x)| < 1 for all x ∈
/ F . Let H = E ∪ F . Then
x∈
/H⇒x∈
/ E, x ∈
/ F ⇒ |f (x)| < ε, |g(x)| < 1 ⇒ |f (x)g(x)| < ε.
Completeness of C0 (X) is left to the reader. This completes the proof.

3.12. Definitions. Let X be a topological space and f : X → K be a
function. The support of f is defined to be the set
supp(f ) = cl{x ∈ X : f (x) 6= 0},
12 1. INTRODUCTION

where cl denotes the closure of a set. A function f is said to have


compact support if supp(f ) is compact.
3.13. Example. For a topological space X let Cc (X) = {f ∈ C(X) :
supp(f ) is compact}. Then (Cc (X), k · k∞ ) is a normed algebra with
pointwise multiplication.
Solution. Let f, g ∈ Cc (X). Then we have
(1) supp(f + g) ⊂ supp(f ) ∪ supp(g).
(2) supp(f g) ⊂ supp(f ).
Now using the facts that union of two compact sets is compact and
every closed subspace of a compact space is compact, it follows that
supp(f + g), supp(f g) are compact. Further, clearly, Cc (X) ⊂ C0 (X).
Hence k · k∞ is submultiplicative on Cc (X). 
3.14. Example. Let P[0, 1] denote the set of all polynomials consid-
ered as functions defined over [0, 1]. Then P[0, 1] is a normed algebra
with usual polynomial multiplication and k · k∞ .
Solution. Since, multiplication of two polynomials is again a poly-
nomial, P[0, 1] is an algebra. Further since P[0, 1] ⊂ C[0, 1], k · k∞
retains its submultiplicativity on P[0, 1] also. 
3.15. Remark. In fact, P[0, 1] can never be a Banach algebra as it has
a countable basis, for example, {1, x, x2 , . . .}. Recall from the course
of Functional Analysis the result we used here.
3.16. Example. Let us write polynomials as an infinite sequences,
i.e., for p(x) = a0 + a1 x + a2 x2 + · · · + an xn , we write p = (a0 , a1 , a2 , . . .)
keeping in mind that ai = 0 for all i > n. Now for two polynomials p =
(a0 , a1 , a2 , . . .) and q = (b0 , b1 , b2 , . . .) define pq = (a0 b0 , a1 b1 , a2 b2 , . . .).
P∞
Also for a polynomial p = (a0 , a1 , a2 , . . .) let kpk1 = |ai |. Then
i=1
P[0, 1] with this product is a normed algebra.
Solution. Left to the reader. 
3.17. Example. For n ∈ N, let C n [0, 1] be the set of all n-times
continuously differentiable K-valued functions defined on [0, 1]. Then
(C n [0, 1], k · k∞ ) is a normed algebra. However, it is not complete in
this norm.
Solution. Left to the reader. 
3. EXAMPLES 13

3.18. Example. Consider C n [0, 1] as in Example 3.17. For x ∈ C n [0, 1],


define n
X 1
kxkn = sup |x(k) (t)|,
i=0
k! 0≤t≤1

where x(k) denotes the k-th derivative of x. Then (C n [0, 1], k · kn ) is a


Banach algebra.
Proof. Left to the reader. 
3.19. Example. Consider the closed unit disc D = {z ∈ C : |z| ≤ 1}.
Let C(D) be the algebra of all C-valued continuous functions defined
on D and A(D) be the set of functions f in C(D) which are analytic on
Int(D). For f ∈ A(D), let kf k∞ = sup |f (z)|. Then A(D) is a Banach
z∈D
algebra, called the disc algebra.
Solution. Requires only the fact that product of two analytic func-
tions is analytic. Prove completeness. 
3.20. Lemma. [KPPP, Theorem 2.26] Let f : [a, b] → R be a func-
tion. Suppose
(1) lim f (x) exists, and is finite for each x0 ∈ (a, b).
x→x0
(2) lim+ f (x) and lim− f (x) also exist, and are finite.
x→a x→b
Then f is bounded on [a, b].
Proof. Left to the readers. 
The set of all K-valued functions defined on [a, b] satisfying the
assumption in the previous lemma is denoted by Lt[a, b]. Using the
above Lemma, we readily see that Lt[a, b] ⊂ B[a, b].
3.21. Conjecture. (Lt[a, b], k · k∞ ) is a Banach algebra.
3.22. Exercise. Prove the following.
(1) For a locally compact Hausdorff space X, Cc (X) is a normed
algebra. Also, Cc (X) ⊂ C0 (X) ⊂ Cb (X). Further, C0 (X) has a
unit if and only if C0 (X) = Cc (X) = C(X) if and only if X is
compact.
(2) (`p , k · kp ), (1 ≤ p ≤ ∞); (B(X), k · k∞ ), (X is a non empty set);
(Cb (X), k · k∞ ), (X is any topological space) and (C0 (X), k · k∞ ),
(X is a locally compact Hausdorff topological space) are complete.
(3) (Cc (R), k · k∞ ) is incomplete.
14 1. INTRODUCTION

(4) (Cc [0, 1], k · k∞ ) is complete. In fact, Cc [0, 1] = C[0, 1].


(5) C0 (N) = c0 ; B(N) = `∞ .
(6) c00 can never be a Banach space.
3.23. Remark. In a polynomial, we use the expressions x, x2 , x3 , . . ..
Let us think of the set S = {x, x2 , x3 , . . .} in which a multiplication
rule is known to us (xs xt = xs+t ). In multiplying two polynomials we
multiply coefficients ai and bj of xi and xj in two polynomials to pro-
duce ai bj xi+j . Then we collect the like terms and add their coefficients.
A further look to the set S reveals that in order to have associativity
of the multiplication of polynomials, we are using the associativity of
multiplication of S. So to define the multiplication of two polynomials,
we are (unknowingly) taking a help of the set S that is closed under
an associative operation. We now use this experience to generate new
examples of algebras.

4. Group Algebras
Let G = {g1 , g2 , . . . , gn } be a finite group. Let A be the set of all
K-valued functions defined on G. An element x of A is denoted by
n
x = Σ x(gi )gi (4.0.1)
i=1

Let us fix x, y ∈ A and a scalar α. Since x, y are functions, clearly,


x = y if x(gi ) = y(gi ) for all i = 1, 2, . . . , n. (4.0.2)
We define
n
x + y = Σ (x(gi ) + y(gi ))gi (4.0.3)
i=1
n
λx = Σ λx(gi )gi (4.0.4)
i=1
and
n
x?y = Σ [ Σ x(gi )y(gj )]gk (4.0.5)
k=1 gi gj =gk

We also define
n
kxk1 = Σ |x(gi )| (4.0.6)
i=1

4.1. Proposition. A defined above is a Banach algebra.


Proof. Treating g1 , g2 , . . . , gn as unknowns we see that elements of
A are linear combinations of n unknowns. Thus a straight verification
shows that A is a normed linear space. Also, {1g1 , 1g2 , . . . , 1gn } forms a
4. GROUP ALGEBRAS 15

basis for A. Thus A is a finite dimensional normed linear space. Hence


A is complete. The definition (4.0.5) shows that A is closed under
the multiplication. Now we fix x = Σni=1 x(gi )gi , y = Σnj=1 y(gj )gj and
z = Σnm=1 z(gm )gm in A and λ ∈ K. Then

n n
(x ? y) ? z = ( Σ x(gi )gi ? Σ y(gj )gj ) ? z
i=1 j=1
n n
={Σ[ Σ x(gi )y(gj )]gk } ? Σ z(gm )gm
k=1 gi gj =gk m=1
n
= Σ{ Σ [ Σ x(gi )y(gj )]z(gm )}gp
p=1 gk gm =gp gi gj =gk
n
= Σ{ Σ x(gi )y(gj )z(gm )}gp . (4.1.1)
p=1 gi gj gm =gp

Also,
n n
x ? (y ? z) = x ? ( Σ y(gj )gj ? Σ z(gm )gm )
j=1 m=1
n n
= Σ x(gi )gi ? { Σ [ Σ y(gj )z(gm )]gk }
i=1 k=1 gj gm =gk
n
= Σ{ Σ x(gi )[ Σ y(gj )z(gm )]}gp
p=1 gi gk =gp gj gm =gk
n
= Σ{ Σ x(gi )y(gj )z(gm )}gp . (4.1.2)
p=1 gi gj gm =gp

Associativity of the multiplication follows from (4.1.1) and (4.1.2). Now


we prove the distributivity of multiplication over addition.
n n n
x ? (y + z) = Σ x(gi )gi ? ( Σ y(gj )gj + Σ z(gj )gj )
i=1 j=1 j=1
n n
= Σ x(gi )gi ? Σ [y(gj ) + z(gj )]gj
i=1 j=1
n
= Σ{ Σ x(gi )[y(gj ) + z(gj )]}gk
k=1 gi gj =gk
n
= Σ{ Σ x(gi )y(gj ) + x(gi )z(gj )}gk
k=1 gi gj =gk
n n
= Σ[ Σ x(gi )y(gj )]gk + Σ [x(gi )z(gj )]gk
k=1 gi gj =gk k=1

= x ? y + x ? z. (4.1.3)
16 1. INTRODUCTION

Left distributivity follows from (4.1.3). Right distributivity can be


proved similarly. Further,
n
λ(x ? y) = Σ [ Σ λx(gi )y(gj )]gk
k=1 gi gj =gk
n n
x ? (λy) = Σ [ Σ x(gi )(λy(gj ))]gk = Σ [ Σ λx(gi )y(gj )]gk
k=1 gi gj =gk k=1 gi gj =gk
n n
(λx) ? y = Σ [ Σ (λx(gi ))y(gj )]gk = Σ [ Σ λx(gi )y(gj )]gk
k=1 gi gj =gk k=1 gi gj =gk

Thus λ(x?y) = x?(λy) = (λx)?y. Now we prove the submultiplicativity


of the norm. For k = 1, 2, . . . , n,
n
x?y = Σ [ Σ x(gi )y(gj )]gk
k=1 gi gj =gk

Hence
n
(x ? y)(gk ) = Σ x(gi )y(gj ) = Σ x(gk gj−1 )y(gj ) (4.1.4)
gi gj =gk j=1

As a result,
n
kx ? yk1 = Σ |(x ? y)(gk )|
k=1
n n
= Σ | Σ x(gk gj−1 )y(gj )|
k=1 j=1
n n
≤ Σ { Σ |x(gk gj−1 )||y(gj )|}
k=1 j=1
n n
= Σ Σ {|x(gk gj−1 )||y(gj )|}
k=1 j=1
n n
= Σ Σ {|x(gk gj−1 )||y(gj )|}
j=1 k=1
n n
= Σ { Σ |x(gk gj−1 )||y(gj )|}
j=1 k=1
n n
= Σ {|y(gj )| Σ |x(gk gj−1 )|}
j=1 k=1
n
= Σ (|y(gj )|kxk1 )
j=1
n
= kxk1 Σ |y(gj )|
j=1

= kxk1 kyk1 .

This completes the proof. 


4. GROUP ALGEBRAS 17

The multiplication defined above in (4.0.5) is called the convolu-


tion and A is called the group algebra of G. It is customary to denote
this algebra by `1 (G). Note that there is yet another multiplication
Σni=1 x(gi )gi Σni=1 y(gi )gi = Σni=1 x(gi )y(gi )gi making A an algebra. But
with this multiplication A behaves just like Kn with pointwise product.
So, we do not investigate A with this multiplication.
What happens when G is an infinite group? The first problem
arises how to express the elements of it’s group algebra. Let us first
assume that G is countable so our notations for series sum carries over.
So let G = {e = g1 , g2 , g3 , . . .} and x : G → K be a function then the

element x of a group algebra would look like x = Σ x(gi )gi . The real
i=1

trouble starts here. Will the series to define the norm kxk1 = Σ |x(gi )|
i=1
converge? It need not. The following proposition takes care of this.

4.2. Proposition. Let G = {e = g1 , g2 , g3 , . . .}. Define


∞ ∞
`1 (G) = {x = Σ x(gi )gi : kxk1 = Σ |x(gi )| < ∞}. (4.2.1)
i=1 i=1

∞ ∞
For x = Σ x(gi )gi , y = Σ x(gi )gi ∈ `1 (G), define
i=1 i=1

∞ ∞ ∞
x?y = Σ [ Σ x(gi )y(gj )]gk = Σ [ Σ x(gk gj−1 )y(gj )]gk . (4.2.2)
k=1 gi gj =gk k=1 j=1

Then (`1 (G), k · k1 ) is a Banach algebra.

Proof. Clearly, `1 (G) is a Banach space. The very first problem we


should sort out is that of convergence of the series within the square
∞ ∞
brackets appearing (4.2.2). Let x = Σ x(gi ), y = Σ y(gj ) ∈ `1 (G).
i=1 j=1

Since Σ |x(gi )| < ∞, the sequence {x(gi )}∞
i=1 is bounded. Suppose
j=1
r = sup |x(gi )|. Then by the Hölder’s inequality,
i∈N


Σ |x(gk gj−1 )y(gj )| ≤ rkyk1 . (4.2.3)
j=1


Thus, for each k, the series Σ x(gk gj−1 )y(gj ) converges in K. Putting
j=1

gk gj−1 = gi , we have Σ x(gk gj−1 )y(gj ) = Σ x(gi )y(gj ). Thus the
j=1 gi gj =gk
18 1. INTRODUCTION

required series converges. As a result, we have for each k ∈ N,



(x ? y)(gk ) = Σ x(gi )y(gj ) = Σ x(gk gj−1 )y(gj ) (4.2.4)
gi gj =gk j=1

Hence,
∞ ∞ ∞
kx ? yk1 = Σ |(x ? y)(gk )| = Σ | Σ x(gk gj−1 )y(gj )|
k=1 k=1 j=1
∞ ∞ ∞ ∞
≤ Σ { Σ |x(gk gj−1 )||y(gj )|} = Σ Σ {|x(gk gj−1 )||y(gj )|}
k=1 j=1 k=1 j=1
∞ ∞ ∞ ∞
= Σ Σ {|x(gk gj−1 )||y(gj )|} = Σ { Σ |x(gk gj−1 )||y(gj )|}
j=1 k=1 j=1 k=1
∞ ∞ ∞
= Σ {|y(gj )| Σ |x(gk gj−1 )|} = Σ (|y(gj )|kxk1 )
j=1 k=1 j=1

= kxk1 Σ |y(gj )| = kxk1 kyk1 .
j=1

Thus x ? y ∈ `1 (G). Rest of the proof is obtained by mimicking the


proof of Proposition 4.1. 

4.3. Proposition. Let G be an at most countable group. Define a map


ϕ : G → `1 (G) by ϕ(gj ) = 1gj , Assume without loss of generality that
g1 is the identity of G. Then the following hold.
(1) ϕ is a one to one map.
(2) ϕ(gk gm ) = ϕ(gk ) ? ϕ(gm ) for all gk , gm ∈ G.
(3) ϕ(g1 ) is identity of `1 (G).
(4) `1 (G) is commutative if and only if so is G.

Proof. (1) follows from the definition of the equality of elements of


`1 (G); see (4.0.2)
(2) ϕ(gk ) ? ϕ(gm ) = 1gk ? 1gm = 1gk gm = ϕ(gk gm ).
(3) For x = Σ x(gi )gi , x ? ϕ(g1 ) = Σ x(gi )gi ? 1g1 = Σ x(gi )gi = x.
i i i

(4) Suppose G is commutative. Let x = Σ x(gi )gi , y = Σ y(gi )gi . For


i i
each k,

x ? y(gk ) = Σ x(gi )y(gj ) = Σ y(gj )x(gi ) = y ? x(gk ). (4.3.1)


gi gj =gk gj gi =gk

Hence x ? y = y ? x. Conversely, suppose gi gj 6= gj gi . Since ϕ is one to


one, 1gi 1gj = ϕ(gi gj ) 6= ϕ(gj gi ) = 1gj 1gi . 
4. GROUP ALGEBRAS 19

Let G = Z = {0, ±1, ±2, . . .}. Here Z is additive group, so our


operation will be + as a result we shall write i + j for gi gj and −i for
gj−1 . Instead of x(gi ) we prefer to write xi . Note that

`1 (G) = `1 (Z) = {x = (xi )∞
−∞ : xi ∈ K; kxk1 = Σ |xi | < ∞}.
i=−∞

For x = (xi )∞ ∞ 1
−∞ , y = (yi )−∞ ∈ ` (Z) we have

x ? y(j) = Σ xi yj−i . (4.3.2)
i=−∞

4.4. Corollary. `1 (Z) is a Banach algebra with convolution.


Recall Example 3.2 of usual `1 . In fact, with the help of the theory
developed so far, we have the following.
4.5. Proposition. For x = (xi ), y = (yj ) ∈ `1 , define

0 if j = 1,
x ? y(j) = j−1 (4.5.1)
 Σ xi yj−i otherwise.
i=1
1
Then (` , k · k) is a Banach algebra.
Proof. The proof is left as a practice work. 
4.6. Exercise. Do as directed.
(1) Prove that `1 does not have identity.
(2) Write x = (xi )∞ 1
−∞ ∈ ` (Z) as

x = Σ xi ti . (4.6.1)
i=−∞

Also, write x = (xi )∞ 1


1 ∈ ` as

x = Σ xi ti . (4.6.2)
i=1

Now multiply respective elements as power series or Laurentz se-


ries. Compare your result with the convolution. Does this help
in showing that `1 does not have identity?
(3) Let Z+ = N ∪ {0}. Define `1 (Z+ ). Show that `1 (Z+ ) with point-
wise addition and scalar multiplication and convolution as multi-
plication is a Banach algebra. Does it have an identity?
(4) Let S be a set with an associative binary operation. Can we define
`1 (S)? If so, will it be an algebra with convolution?
(5) For (x1 , y1 ), (x2 , y2 ) ∈ K2 , define the following.
20 1. INTRODUCTION

(a) (x1 , y1 ) 1 (x2 , y2 ) = (x1 y1 , 0).


(b) (x1 , y1 ) 2 (x2 , y2 ) = (x2 y2 , 0).
Show that K2 is a Banach algebra with each of 1 and 2 .

5. New From Old


In this section we construct new algebras out of the given algebras.
5.1. Definitions. Let A be a Banach algebra and B ⊂ A. B is called
a subalgebra of A if B is also an algebra with the operations inherited
from A. Further, if A has identity and if B contains the identity of A,
then B is called the unital subalgebra of A. Let I be a subalgebra of
A. I is called a left ideal of A if xI ⊂ I for every x ∈ A. I is called a
right ideal of A if Ix ⊂ I for every x ∈ A. I is said to be a two sided
ideal (or for brevity, an ideal ), if it is a left as well as right ideal of A.
An ideal of A is said to be proper I = 6 A.
5.2. Remark. All ideals of a commutative algebra are two sided ideals.
Before reading the following theorem, the reader is urged to recall
the topic “completion of a metric space.”
5.3. Theorem. Let (A, k · k) be a normed algebra. Let à denote the
metric space completion of A. For x, y ∈ Ã and λ ∈ K, define the
operations and norm on à in the following way.
(1) Choose sequences {xn }, {yn } ∈ A such that xn → x and yn → y.
(2) Define x + y = lim xn + yn .
n→∞
(3) Define λx = lim λxn .
n→∞
(4) Define xy = lim xn yn .
n→∞
(5) Define kxk = lim kxn k.
n→∞

Then with these operations and the norm à becomes a Banach algebra
and (A, k · k) becomes a subalgebra of Ã.
Proof. First we show that the sequences {} 
5.4. Example. The following lists several examples of subalgebras and
ideals of different algebras. The proofs are straight verifications left to
the reader.
(1) Let A0 = (C[a, b], k · k∞ ) and An = (C n [a, b], k · k∞ ). Then each
An is a subalgebra of An−1 , n ∈ N. Also let x ∈ [a, b] be fixed.
5. NEW FROM OLD 21

Define
Inx = {f ∈ An : f (x) = 0}.
Then Inx is an ideal of An .
(2) Let A be an algebra and x ∈ A. Then
Ax = {ax : a ∈ A}
is a left ideal of A.
(3) Consider Kn with pointwise operations. For 1 ≤ k ≤ n, define
Ik = {x = (xi ) ∈ Kn : xk = 0}. Then Ik is an ideal of Kn .
(4) c00 is an ideal of `p , (1 ≤ p ≤ ∞).
(5) c00 is an ideal of c as well as of c0 .
(6) c0 is an ideal of c.
n
(7) I = {p = a0 + a1 x + a2 x2 + · · · + an xn ∈ P[0, 1] : Σ ai = 0} is
i=1
and ideal of P[0, 1].
n
(8) I = {p = a0 + a1 x + a2 x2 + · · · + an xn ∈ P[0, 1] : Σ (−1)i ai = 0}
i=1
is and ideal of P[0, 1].
5.5. Theorem. Let A be an algebra with identity. Let x ∈ A. Then
the following are equivalent.
(1) x has a left inverse.
(2) x does not belong to any left ideal of A.
Proof. Assume (1). Let y be a left inverse of x. Suppose I is a proper
left ideal of A such that x ∈ I. Then 1 = yx ∈ I. Hence for any z ∈ A,
we have z = z1 = (zy)x ∈ I, giving A ⊂ I, a contradiction. Thus (2)
follows.
Assume (2). Clearly, x = 1x ∈ Ax and Ax is a left ideal of A. So, Ax
cannot be a proper ideal. That is, A = Ax. So, 1 ∈ Ax. As a result,
1 = yx for some y ∈ A. Thus x is left regular. 
The following theorem can be proved similarly.
5.6. Theorem. Let A be an algebra with identity. Let x ∈ A. Then
the following are equivalent.
(1) x has a right inverse.
(2) x does not belong to any right ideal of A.
5.7. Corollary. A proper left or right ideal I of an algebra A cannot
contain any invertible element of A. In particular, it cannot contain 1.
22 1. INTRODUCTION

5.8. Example. Let A = K×K with pointwise addition and scalar mul-
tiplication. For (x1 , y1 ), (x2 , y2 ) ∈ K2 define (x1 , y1 )(x2 , y2 ) = (x1 x2 +
x1 y2 + x2 y1 , y1 y2 ) and k(x1 , y1 )k1 = |x1 | + |y1 |. Then (A, k · k1 ) is a
Banach algebra. Now let B = {(x, 0) : x ∈ K}. Then the following are
easily seen.
(1) A has unit element (0, 1).
(2) B has unit element (1, 0).
Thus A and B are both unital algebras and B is a subalgebra of A.
However, since unit of A is not in B, B is not a unital subalgebra of A.
Also note here that B is an ideal of A and B has identity. But since it
is proper, it cannot contain the identity of A.

5.9. Example. Let A = K×K with pointwise addition and scalar mul-
tiplication. For (x1 , y1 ), (x2 , y2 ) ∈ K2 define (x1 , y1 )(x2 , y2 ) = (x1 x2 , 0)
and k(x1 , y1 )k1 = |x1 | + |y1 |. Then (A, k · k1 ) is a Banach algebra. Now
let B = {(x, 0) : x ∈ K}. Then the following are easily seen.
(1) A does not have a unit element.
(2) B has unit element (1, 0).
Thus A is not a unital algebra but B is a unital algebra. Also, B is a
subalgebra of A. However, B is not a unital subalgebra of A. Why?

5.10. Example.  Let M2 denote the algebra of all 2 × 2 matrices over
a a
K. Define B = : a ∈ K . Then B is a subalgebra of M2 .
a a
Here M2 and B both are unital but B is not a unital subalgebra of
M2 .
  
ka ma
5.11. Exercise. For fixed k, m ∈ K, let C = :a∈K .
ka ma
Show that C is a subalgebra of the algebra M2 . Obtain conditions on
k, m so that the algebra C becomes a unital algebra. In any case, C is
not a unital subalgebra of M2 .

5.12. Example. Let (Ai , k · ki ), (1 ≤ i ≤ n) be normed algebras and


n
A denote the cartesian product Π Ai with pointwise operations. Also
i=1
n
for x = (xi ) ∈ A, define kxk = Σ kxi ki . Then (A, k · k) is a normed
i=1
algebra. Also, A is complete if and only if each Ai is complete.
6. HOMOMORPHISMS 23

6. Homomorphisms
6.1. Definitions. Let A, B be two algebras over the same field K. A
linear mapping ϕ : A → B is said to be an algebra homomorphism from
A to B if ϕ(xy) = ϕ(x)ϕ(y), for all x, y ∈ A. The set ker(ϕ) = {x ∈
A : ϕ(x) = 0} is called the kernel of ϕ. Further, if ϕ is one-one, then
it is called an algebra isomorphism. If ϕ is one-one and onto, then it
is called an onto algebra isomorphism and in this case, we say that A
is algebraically isomorphic to B. Further, suppose that A and B are
normed algebras. If the algebra isomorphism ϕ is a homeomorphism
between, A and ϕ(A), then we say that ϕ is a topological isomorphism
and A and ϕ(A) are topologically isomorphic. Further if kϕ(x)k = kxk
for all x ∈ A, then ϕ is called an isometric isomorphism and A and
ϕ(A) are called isometric isomorphic.

6.2. Remark. While dealing with more than one algebras, we shall
assume that they are over a same field K, unless otherwise stated.

6.3. Proposition. Let I be an ideal of a normed algebra A. Define


A/I = {x + I : x ∈ A}. For x, y ∈ A and α ∈ K, define,
x+I +y+I =x+y+I
(x + I)(y + I) = xy + I
α(y + I) = αy + I.
Then A/I is an algebra. Further, suppose that I is closed in A. Then
the following hold.
(1) k|x + I|k = inf{kx + yk : y ∈ I} is an algebra norm on A/I.
(2) The canonical map π : A → A/I defined by π(x) = x+I, (x ∈ A)
is a continuous open algebra homomorphism with ker(π) = I.
(3) A is a Banach algebra if and only if A/I is a Banach algebra.

Proof. You know from Functional Analysis [Lim] the following.


(i) A/I is a linear space.
(ii) If I is closed in A, then A/I is a normed linear space.
(iii) In the second case, A is complete if and only if A/I is complete.
(iv) The canonical map π is continuous and open.
For x1 + I = x2 + I and y1 + I = y2 + I ∈ A/I,
x1 y1 − x2 y2 = x1 (y1 − y2 ) + (x1 − x2 )y2 ∈ I ⇒ x1 y1 + I = x2 y2 + I.
24 1. INTRODUCTION

Thus the multiplication is well defined on A/I. The associativity,


distributivity etc. are routine verifications. Now let x+I, y +I ∈ A/I.
Let ε > 0. Find a, b ∈ I such that kx + ak < k|x + I|k + ε and
ky + bk < k|y + I|k + ε. Then ay + xb + ab ∈ I. Consequently,
k|xy + I|k ≤ kxy + ay + xb + abk
= k(x + a)(y + b)k
≤ kx + akky + bk
< (k|x + I|k + ε)(k|x + I|k + ε).
Taking infimum over all ε > 0 on the right hand side, we get,
k|xy + I|k ≤ k|x + I|kk|y + I|k.
This completes the proof. 
6.4. Proposition. Let A and B be two algebras and ϕ : A → B be a
homomorphism. Suppose A1 is a subalgebra of A and I is an ideal of
A. Also suppose that B1 is a subalgebra of B and J be an ideal of B.
Then
(1) ϕ(A1 ) is a subalgebra of B. In particular, ϕ(A) is a subalgebra of B.
(2) ϕ(I) is an ideal of ϕ(A).
(3) ϕ−1 (B1 ) is a subalgebra of A.
(4) ϕ−1 (J ) is an ideal of A.
(5) ker(ϕ) is an ideal of A. Also, ϕ is an isomorphism if and only if
ker(ϕ) = {0}.
(6) Define ϕ̃ : A/ ker(ϕ) → B by ϕ̃(x + ker ϕ) = ϕ(x). Then ϕ̃ is
an isomorphism. Further if A, B are normed algebra and if ϕ is
continuous, then ϕ̃ is continuous.
Proof. Proof is a routine verification you have done for many similar
structures. 

7. Adjoining Identity
In general, an algebra need not have identity. In that case, we can
obtain a bigger algebra with identity containing the given algebra as a
subalgebra.
7.1. Theorem. [La, p.10] Let A be a normed algebra without identity.
Define Ae = A × K. For (x, λ), (y, µ) ∈ Ae and k ∈ K, define
(x, λ) + (y, µ) = (x + y, λ + µ).
7. ADJOINING IDENTITY 25

k(x, λ) = (kx, kλ).


(x, λ)(y, µ) = (xy + λy + µx, λµ). (7.1.1)
Also, for (x, λ) ∈ Ae , let k(x, λ)k1 = kxk + |λ|. Then
(1) (Ae , k · k) is a unital normed algebra.
(2) ι : A → Ae , defined by ι(x) = (x, 0), (x ∈ A) is an isometric
isomorphism.
(3) Identifying A with {(x, 0) : x ∈ A}, A is a closed ideal of Ae .
(4) Ae is complete if and only if A is complete.
Proof. (1) Clearly, (Ae , k · k1 ) is a normed linear space. Now let
(x, λ), (y, µ), (z, θ) ∈ Ae Then
(x, λ)((y, µ)(z, θ)) = (x, λ)(yz + µz + θy, µθ)
= (x(yz + µz + θy) + λ(yz + µz + θy) + µθx, λ(µθ))
= ((xy)z + µxz + θxy + λyz + λµz + λθy + µθx, (λµ)θ)
= ((xy + λy + µx)z + θ(xy + λy + µx) + λµz, (λµ)θ)
= (xy + λy + µx, λµ)(z, θ)
= ((x, λ)(y, µ))(z, θ).
This proves that the multiplication is associative. Also,
(x, λ)((y, µ) + (z, θ)) = (x, λ)(y + z, µ + θ)
= (x(y + z) + (µ + θ)x + λ(y + z), λµ + λθ)
= (xy + xz + µx + θx + λy + λz, λµ + λθ)
= (xy + µx + +λy+, λµ) + (xz + θx + λz, λθ)
= (x, λ)(y, µ) + (x, λ)(z, θ).
Similarly, we can prove ((y, µ)+(z, θ))(x, λ) = (y, µ)(x, λ)+(z, θ)(x, λ).
We leave the proof of
α((x, λ)(y, µ)) = (α(x, λ))(y, µ) = (x, λ)α(y, µ)
to the reader. The identity (x, λ)(0, 1) = (x, λ) = (0, 1)(x, λ) is a
straight computation. (compute!). Finally,
k(x, λ)(y, µ)k1 = k(xy + λy + µx, λµ)k1
= kxy + λy + µxk + |λµ|
≤ kxyk + kλyk + kµxk + |λµ|
≤ kxkkyk + |λ|kyk + |µ|kxk + |λ||µ|
26 1. INTRODUCTION

= kxk(kyk + |µ|) + |λ|(kyk + |µ|)


= (kxk + |λ|)(kyk + |µ|)
= k(x, λ)k1 k(y, µ)k1

(2) Clearly, ι(x + y) = (x + y, 0) = (x, 0) + (y, 0) = ι(x) + ι(y)


and ι(kx) = (kx, 0) = k(x, 0) = kι(x). Also, ι(xy) = (xy, 0) =
(x, 0)(y, 0) = ι(x)ι(y). Also, kι(x)k = k(x, 0)k1 = kxk, showing that ι
an isometric isomorphism.
(3) Let (xλ) ∈ Ae and (y, 0) ∈ A. Then (x, λ)(y, 0) = (xy + λy, 0) ∈ A
and (y, 0)(x, λ) = (yx, λy, 0) ∈ A. Thus A is an ideal of Ae . Now let
{(xn , 0)} be a sequence in A such that {(xn , 0)} → (x, λ). Then
|λ| ≤ kxn − xk + |λ| = k(xn , 0) − (x, λ)k1 → 0
Thus λ = 0. Hence, (x, λ) = (x, 0) ∈ A, proving that A is closed in
Ae .
(4) If Ae is complete, then by (3) A is complete. Conversely, suppose
A is complete. Let {(xn , λn )} be a Cauchy sequence in Ae . Then
|λn − λm | ≤ kxn − xm k + |λn − λm |
= k(xn , λn ) − (xm , λm )k1 → 0 as n, m → ∞.
Thus {λn } is Cauchy in K. Since K is complete, λn → λ for some
λ ∈ K. Also,
kxn − xm k ≤ kxn − xm k + |λn − λm |
= k(xn , λn ) − (xm , λm )k → 0 as n, m → ∞.
Thus {xn } is Cauchy in A. Since A is complete, xn → x for some
x ∈ A. But then
k(xn , λn ) − (x, λ)k1 = kxn − xk + |λn − λ| → 0 as n → ∞,
proving that (xn , λn ) → (x, λ). Thus Ae is complete. 

8. Regular Elements
The set of all regular elements play an important role in the theory
of Banach algebras.
8.1. Theorem. Let A be a Banach algebra with identity. Then the
following hold
8. REGULAR ELEMENTS 27

(1) G is a group.

(2) If x ∈ A and if k1 − xk < 1, then x ∈ G and x−1 = 1+ Σ (1−x)n .
n=1
(3) G is open in A.
(4) The mapping x ∈ G 7→ x−1 ∈ G is a homeomorphism.

Proof. (1) For x, y ∈ G, (xy)−1 = y −1 x−1 . So, xy ∈ G. Associativity


of the multiplication is by definition. Also, 1 ∈ G. Finally, x ∈ G ⇒
(x−1 )−1 = x ⇒ x−1 ∈ G.
∞ ∞
(2) Suppose k1 − xk < 1. So, Σ k(1 − x)n k ≤ Σ k1 − xkn < ∞.
n=1 n=1

n
Hence the series Σ k(1 − x) k is absolutely convergent series in the
n=1

Banach algebra A. As a result, Σ (1 − x)n converges in A. Let y =
n=1

n
Σ (1 − x) . Then
n=0

y − xy = (1 − x)y

= (1 − x) Σ (1 − x)n
n=0

= Σ (1 − x)n
n=1

= 1 + Σ (1 − x)n − 1
n=1

= Σ (1 − x)n − 1
n=0
=y−1

Thus xy = 1. Also,

y − yx = y(1 − x)

= Σ (1 − x)n (1 − x)
n=0

= Σ (1 − x)n
n=1

= 1 + Σ (1 − x)n − 1
n=1

= Σ (1 − x)n − 1
n=0
=y−1

Thus yx = 1.
28 1. INTRODUCTION

−1
(3) Let x0 ∈ G and kx − x0 k < kx−1
0 k . Then
−1
k1 − x−1 −1 −1 −1 −1
0 xk = kx0 (x0 − x)k ≤ kx0 kk(x0 − x)k < kx0 kkx0 k = 1.
Hence x−1 −1 −1
0 x ∈ G. Thus x0 , x0 x ∈ G. So, x = x0 (x0 x) ∈ G.
(4) Let x0 ∈ G, we show that(the function x 7→)x−1 is continuous at
ε 1
x0 . For an ε > 0, let δ = min ,
−1 2 2kx−1 k
. Let kx − x0 k < δ.
2kx0 k 0
Then
1 1 1
kx−1 −1 −1 −1
0 x − 1k = kx0 (x − x0 )k ≤ kx0 kkx − x0 k < kx0 k −1 = = 1− .
2kx0 k 2 2
Hence,
1 − kx−1
0 x − 1k > 1/2.
So,
1
< 2. (8.1.1)
1− kx−1
0 x − 1k

Consequently, x−1 x0 = (x−1
0 x)
−1
= Σ (1 − x−1 n
0 x) . Thus
n=0

kx −1
− x−1
0 k = k(x−1
0 x − 1)x−1
0 k

≤ kx−1 −1
0 x − 1kkx0 k

= kx−1 −1 n
0 kk Σ (1 − x0 x) − 1k
n=0

≤ kx−1 −1 n
0 kk Σ (1 − x0 x) k
n=1
∞ n
≤ kx−1 −1
0 k Σ k1 − x0 xk
n=1
−1
kx0 kk1 −x−1
0 xk
= −1
1 − k(1 − x0 x)k
= 2kx−1 −1
0 kkx0 (x0 − x)k
2
≤ 2kx−1
0 k kx0 − xk
2 2 ε
< 2kx−1 −1
0 k δ ≤ 2kx0 k 2 < ε.
2kx−1
0 k

This proves that x 7→ x−1 is continuous at x0 ; and so on G. Since the


mapping is the inverse of itself, it is a homeomorphism. 
In the absence of identity of A, it is a common practice to define
the concept of quasi-invertibility or quasi-regularity. We shall need this
concept to investigate algebras without unit.
9. TOPOLOGICAL DIVISORS OF ZERO 29

8.2. Definitions. Let A be an algebra. For x, y ∈ A, we define a new


product ◦ on A by x ◦ y = x + y − xy. An element x ∈ A is said to
be left quasi-invertible or left quasi-regular if there exists y ∈ A such
that y ◦ x = 0. In this case, y is called right quasi-invertible or right
quasi-regular. Also, y is called a left quasi-inverse of x and x is called
a right quasi-inverse of y. If x has a left and a right inverse, then x is
said to be quasi-regular or quasi-invertible.

9. Topological Divisors of Zero


In a ring R, the concept of zero divisor well investigated. However,
in the presence of topology we shall topologize this concept.

9.1. Definitions. Let A be a normed algebra and x be a nonzero


element of A. x is said to be a zero divisor or a divisor of zero if there
is y 6= 0 such that xy = 0 or yx = 0. If there is a sequence {xn } in
A such that kzn k = 1 such that either xxn → 0 or xn x → 0, then x is
said to be a topological divisor of zero. The set of topological divisors
of A is denoted by Z.

9.2. Theorem. Let A be a Banach algebra with unit. Then Z ⊂ S.


Further, the boundary of S is subset of Z.

Proof. Let x ∈ A be a topological divisor of zero. Let {xn } be a


sequence in A such that kxn k = 1 such that xxn → 0. Suppose, if
possible, that z is invertible. Then because of the (joint) continuity
of multiplication, xn = x−1 xxn → 0. So, kxn k → 0, contradiction.
Thus x cannot be invertible. Similarly, if {xn } is a sequence in A such
that kxn k = 1 such that xn x → 0, then xn = xn xx−1 → 0, again a
contradiction. Thus x cannot be invertible.
Now suppose that x is a boundary point of S. Then there ex-
ists a sequence {xn } from G, such that xn → x. Suppose, if possi-
ble, that {x−1 −1 −1
n } is bounded. But then k1 − xn xk = kxn (xn − x)k ≤
kx−1 −1
n kkxn − xk → 0. Thus for some m ∈ N, k1 − xm xk ≤ 1. Hence,
xm x is invertible. Consequently, x (as a product (xn )(x−1
−1
n x) of two
invertible elements) is invertible. This contradicts the singularity of
x. Thus {x−1n } is unbounded. Hence there exists a subsequence {xnk }
x−1
nk
of {xn } such that kxnk k > k, (∀ k = 1, 2, 3, . . .). Define zk = .
kx−1
nk k
30 1. INTRODUCTION

Then kzk k = 1, ∀ k. Also,


xx−1
nk 1 + xx−1
nk − 1 1 + xx−1 −1
nk − xnk xnk
xzk = = =
kx−1
nk k kx−1
nk k kx−1
nk k

1 xx−1 −1
nk − xnk xnk 1 (x − xnk )x−1
nk
= + = +
kx−1
nk k kx −1
nk k kx −1
nk k kx −1
nk k
1
= −1 + (x − xnk )zk → 0
kxnk k
Thus x is a topological divisor of zero. 
9.3. Corollary. Let x be a singular element in a Banach algebra A
and {xn } be a sequence of invertible elements such that xn → x. Then
{x−1
n } is unbounded.

Let A be a Banach algebra with identity, B a unital subalgebra of


A and x ∈ B. If x is regular in B, then x is regular in A also. However,
if x is singular in B, then it need not remain singular in A. This can
be seen by the following example.
9.4. Example. Consider the Disc algebra (see Example 3.19). Let
Γ = {λ ∈ C : |λ| = 1} denote the unit circle. By restricting the domain
of f ∈ A(D) to Γ, we see that A(D) is a subalgebra of C(Γ). Then
1
f (z) = z is singular in A(D) as cannot be defined on D. However,
z
1
g(z) = is defined and continuous on Γ. Thus f is regular in C(Γ).
z
9.5. Definition. Let B be a Banach algebra with identity. A singular
element x ∈ B is called permanently singular if x remains singular in
A whenever B is a unital subalgebra of A.
Topological divisors of zero are not only singular but they are
permanently singular.
9.6. Proposition. Let A be a Banach algebra with unit and B be its
subalgebra. If x ∈ B is a topological divisor of zero, then z is singular
in A.
Proof. Without loss of generality we take {xn }, a sequence in B such
that kxn k = 1 and xxn → 0. Then {xn } is a sequence in A such that
kxn k = 1 and xxn → 0. Thus x is a topological divisor of zero in A.
So, x is singular in A. 
10. SPECTRUM 31

10. Spectrum
As we know the eigenvalues of a square matrix plays important
role in investigating the matrix. Let x be an n × n matrix and e denote
the n×n matrix. Then λ is an eigenvalue of x iff x−λe is not invertible.
We generalize this concept to the setup of an algebra.
10.1. Definition. Let A be an algebra with identity 1. For x ∈ A, we
define the spectrum of x to be the set
spA (x) = {λ ∈ K : x − λ1 is not invertible in A}.
It is also denoted by σA (x). The complement of the spectrum of x is
called the resolvent set of x and is denoted by ρA (x).
Note here that the computation of the spectrum of an element
depends on the algebra. As earlier, if we consider the function f (z) = z,
then f is not invertible in A(D), but is invertible in C(Γ). Thus 0
belongs to the spectrum of f calculated with respect to A(D) but 0
does not belong to the spectrum of f calculated with respect to C(Γ).
However, we will drop A from the notation spA (x) and ρA (x), whenever
there is no room for confusion.
10.2. Lemma. Let A be an algebra with identity 1 and x ∈ A. Then
sp(x) is closed and bounded subset of K. That is, sp(x) is a compact
subset of x.
Proof. Let λ ∈ K such that |λ| > kxk. Then k λx k < 1. Hence, 1 − λx
is invertible. Consequently, (x − λ1) = (−gl)(1 − λx ) is invertible. That
is λ ∈ sp(x) ⇒ λ ≤ kxk. Also, consider the continuous map defined as
λ ∈ K 7→ x − λ1 ∈ A. Then sp(x) is the inverse image of the closed
set A \ G. Thus sp(x) must be closed. 
10.3. Remark.
(1) The completeness of A cannot be dropped from the above Lemma.
Indeed, for any nonconstant polynomial p ∈ P[0, 1], sp(x) = K.
(2) In general, for a real Banach algebra A, spectrums of some of its
elements may be empty. Indeed, considering C as a real algebra,
we see that the invertible elements are precisely, the non-zero
elements. But i − λ 6= 0 for any λ ∈ R. Hence the spectrum of
i is empty. However, the complex algebras are very nice! They
assure that the spectrums of their elements are nonempty.
32 1. INTRODUCTION

Let A be a unital Banach algebra and x ∈ A. For λ ∈ ρ(x), define


x(λ) = (x−λ1)−1 . This map is called the resolvent of x. Let us observe
that resolvent is the composition of the following maps.
λ ∈ ρ(x) 7→ λ1 7→ x − λ1 ∈ G 7→ (x − λ1)−1 ∈ G.
Thus the resolvent of x is a continuous mapping from ρ(x) → A. Note
also that x(λ) = (x − λ1)−1 = λ1 ( λx − 1)−1 . Taking λ → ∞, we see that
1
λ
→ 0 and λx − 1 → −1. As a result,
lim x(λ) = 0. (10.3.1)
λ→∞
Further,
x(λ) = x(λ)(x − µ1)(x − µ1)−1
= x(λ)(x − λ1 + λ1 − µ1)(x − µ1)−1
= x(λ)(x − λ1 + λ1 − µ1)x(µ)
= x(λ)((x − λ1) + (λ1 − µ1))x(µ)
= (x(λ)(x − λ1) + x(λ)(λ1 − µ1))x(µ)
= (1 + x(λ)(λ1 − µ1))x(µ)
= x(µ) + x(λ)(λ1 − µ1)x(µ)
Consequently,
x(λ) − x(µ)
= x(λ)x(µ) (10.3.2)
λ−µ
Equation (10.3.2) is known as the resolvent equation. Before we
proceed further to prove that spectrum of each element of a complex
Banach algebra is nonempty, we need to mention some of the result
that you have learnt in the introductory course of Functional Analysis.
10.4. Lemma. Let X be a normed linear space and x ∈ X be nonzero.
Then there exists a continuous linear map f : X → K such that f (x) 6=
0. In other words, if f (x) = 0 for all continuous linear functional f on
X, then x = 0.
10.5. Theorem. Let A be a complex Banach algebra with identity and
x ∈ A. Then spA (x) 6= ∅.
Proof. For a continuous linear functional f on A, let f : ρ(x) → C be
defined by f (λ) = f (x(λ)). Then from the resolvent equation (10.3.2),
f (λ) − f (µ)
lim = f (µ)2 .
λ→µ λ−µ
11. GEL’FAND MAZUR THEOREMS 33

Thus f is differentiable on whole of ρ(x). Now assume, if possible,


that sp(x) = ∅. Then f is an entire function. On the other hand,
(10.3.1) and continuity of f show that f (λ) → 0 as λ → ∞. That
is, f is an entire bounded function. So, f must be constant. Since,
f (λ) → 0 as λ → ∞, we must have f (x(λ)) = f (λ) = 0 for all λ ∈ C.
Since, f is an arbitrary continuous linear functional on A, by Lemma
10.4, (x − λ1)−1 = x(λ) = 0, a contradiction. Thus sp(x) cannot be
empty. 

11. Gel’fand Mazur Theorems


The Gel’fand Mazur Theorem asserts that every complex Banach
division algebra is topologically isomorphic to C. Recall that a division
algebra is an algebra in which every nonzero element is invertible. These
type of algebras are always assumed to be unital.
11.1. Theorem (Gel’fand Mazur Theorem). Let A be a division alge-
bra over C. If A is a division algebra, then A is isomorphic to C.
Proof. Let A be a Banach division algebra and x ∈ A. Suppose, if
possible, that x−λ1 6= 0 for any λ ∈ C. But then x−λ1 is invertible for
every λ ∈ C. This means that sp(x) is an empty set, a contradiction.
Thus there exists λ ∈ C such that x − λ1 = 0, that is, x = λ1.
Thus A = {λ1 : λ ∈ C}. Consequently, λ1 ∈ A 7→ λ is an algebra
isomorphism. Since every linear map on a finite dimensional normed
linear space is continuous, we assert that this map is also a topological
isomorphism. This completes the proof. 
This is the result proved by Gel’fand and then improved by Mazur.
Later on many such characterizations of C are proved. All of them are
known as the Gel’fand Mazur Theorems. In fact, in the general setup,
one starts with an algebra A over K where K is either R or C. Then
with some conditions on it, it is proved that A = br, C or H. H is the
noncommutative division algebra of all quaternions.
11.2. Lemma. Let A be a unital Banach algebra and x ∈ A. Let λ be
a boundary point of sp(x). Then x − λ1 is either 0 or a boundary point
of S. In particular, in this latter case, x is topological divisor of zero.
Proof. Suppose x − λ1 6= 0. Since λ is a boundary point of sp(x),
there exists a sequence {λn } from ρ(x) such that λn → λ. But then
34 1. INTRODUCTION

x − λn 1 → x − λ1. Since λn ∈ ρ(x), x − λn 1 ∈ G. Thus x − λ is a


boundary point of S. Hence x − λ is a topological divisor of zero. 

Before we go further, we state one result of topology.

11.3. Lemma. Let X be a connected topological space and Y be a


nonempty proper subset of X, then Y has at least one boundary point.

11.4. Theorem. Let A be a Banach algebra over C. Suppose any one


of the following holds.
(1) A has no topological divisor of zero.
(2) For some M > 0, kxyk ≥ M kxkkyk for all x, y ∈ A.
(3) For x ∈ G, kx−1 k ≤ 1 .
kxk
Then A is isomorphic to C.

Proof. (1) Let x be a nonzero element of A. First note that sp(x)


is a nonempty compact subset of C. Thus sp(x) is nonempty and
proper. So, by the previous lemma, we assert that sp(x) has at least
one boundary point say, λ. But then x − λ is either 0 or a topological
divisor of zero. Since A has no topological divisor of zero, we conclude
that x − λ1 = 0. But then x = λ1 is invertible. So, A is a division
algebra. Hence A is isomorphic to C.
(2). Suppose if possible x ∈ A is a topological divisor of zero. Then
there exists a sequence {xn } in A such that kxn k = 1 and xn x → x.
But then kxn kkxk ≤ Kkxn xk → 0. Since kxn k = 1, we conclude that
x = 0. Thus A does not have any (nonzero) topological divisor of zero.
Hence A is isomorphic to C.
(3). Let 0 6= x ∈ A. Let λ be a boundary point of sp(x). If possible,
suppose that x 6= λ1. Choose a sequence xn in G such that xn → x − λ.
Hence, kxn k → kx − λ1k. Since, kx − λ1k 6= 0, 1 → 1
.
kxn k kx − λ1k
Thus 1 is a bounded sequence. As a result, kx−1 1
n k ≤ kx k gives
kxn k n
boundedness of kx−1n k. This is a contradiction to Corollary 9.3. Thus
x = λ1. 

12. Algebras Without Identity


This section gives some glimpses as to how the theory of nonunital
algebras could be developed.
12. ALGEBRAS WITHOUT IDENTITY 35

12.1. Remark. For an algebra A, we shall write elements of Ae as


x + λe, where x ∈ A and λ ∈ K. Also, the elements of A will be
denoted by x instead of x + 0e.
12.2. Theorem. Let A be an algebra and Ae be its unitization, i.e., the
algebra obtained by adjoining the identity 1 to A. Then the following
hold. Let x, y ∈ A.
(1) x is left(respectively, right) quasi-invertible in A if and only if
e − x is left(respectively, right) invertible in A. In this case, if y
is a left (respectively, right) quasi-inverse of x in A, then e − y is
a left (respectively, right) inverse of e − x in A.
(2) If x is quasi-regular, then its quasi-inverse is unique.
Proof. (1) Clearly, x ◦ y = 0 ⇔ x + y − xy = 0 ⇔ (e − x)(e − y) =
−x − y + xy + e = e.
(2) Suppose x has two quasi-inverses, y1 , y2 . Then by (1), (e − y1 ), (e −
y2 ) are two inverses of (e − x) in Ae , a contradiction to the uniqueness
of inverses in any ring. Thus x can have at most one quasi-inverse. 
12.3. Remark. We note that in general, an element can have more
than one left (respectively, right) quasi-inverses. However, if it has
one left and one right quasi-inverse, then they are same. Also, it is a
common practice to denote the quasi-inverse of x by x−1 .
12.4. Proposition. Let A be an algebra. Fix x ∈ A and ci = {−z +
zx : z ∈ A} is a left ideal of A. Further, x has a left quasi-inverse in
A if and only if I = A.
Proof. Clearly, {−z+zx : z ∈ A} is a left ideal of A. Further suppose
x is left quasi-regular and y ◦ x = 0. Then x = −y + yx ∈ A. As a
result, for any z ∈ A, z = (z − zx) + zx ∈ I + I = I. Thus I = A.
Conversely, suppose I = A. Then x ∈ I. So, there exists a y ∈ A such
that x = −y + yx, i.e., x + y − yx = 0. 
12.5. Exercise.
(1) Obtain the analogue of Proposition 12.4 for right in place of left.
12.6. Definition. Let A be a nonunital Banach algebra. Define the
spectrum of x ∈ A to be the spectrum of (x, 0) ∈ Ae .
12.7. Remark. Refer to [La] for the theory of nonunital Banach al-
gebras. We stop the discussion here as it is beyond the scope of our
syllabus.
Bibliography

[Ban] S. Banach, Théorie des Opŕations linéaires, Monografje Matematyczne, Vol.


1, Warsaw, 1932.
[B] S.K. Berberian, Introduction to Hilbert space, Oxford University Press, New
York, 1961.
[BD] F.F. Bonsall and J. Duncan, Complete Normed Algebras, Springer-Verlag,
1973.
[Br] A. Browder, Introduction to Function Algebras, W.A. Benjamin, Inc., New
York, 1969.
[DS] N. Dunford and J.T. Schwartz, Linear Operators, Interscience Publishers, A
division of John Wiley & Sons, Inc., New Yourk, Part I, 1958 Part II 1963.
[GS] I.M. Gelfand, D. Raikov and G.E. Shilov, Commutative Normed Rings,
Chelsea Publishing Company, New York, 1964 (Original Russian edition 1960).
[H] P.R. Halmos, A Hilbert Space Problem Book, D. Van Nostrand Company, Inc.,
Princeton, New Jersey, 1967.
[KPPP] D.J. Karia, M.L. Patel, N.Y. Patel and B.P. Patel, A Textbook of Calculus
with an Introduction to Differential Equations, Roopal’s Mathematics Series
1, Roopal Prakashan, Vallabh Vidyanagar, 2003.
[La] R. Larsen, Banach Algebra, Marcel Dekker Inc. New York, 1973.
[Lim] B.V. Limaye, Functional Analysis, (Second Edition), New Age International
Ltd., New Delhi, 1996.
[N] M.A. Naimark, Normed Rings, Erven P.Noordhoff Ltd., Groningen, the
Netherlands, 1960 (Original Russian edition 1955).
[Ri] C.E. Rickart, General Theory of Banach Algebras, D. Van Nostrand Company,
Inc., Princeton, New Jersey, 1960.
[RR] H. Radjavi and P. Rosenthal, Invariant subspaces, Springer Verlag, New York,
1973.
[Ru1] W. Rudin, Real and Complex Analysis, McGraw-Hill Book Company Inc.
New York, 1966.
[Ru2] W. Rudin, Functional Analysis, (TMH Edition 7th Printing) Tata McGraw-
Hill Publishing Co. Ltd., New Delhi, 1982. (originally published by McGraw
Hill, Inc., 1973)
[S] G.F. Simmons, Introduction to Topology and Modern Analysis, McGraw-Hill
Book Company Inc., New York, 1963.
[St] M.H. Stone, Linear Transformations in Hilbert Space and Their Applications
to Analysis, Amer. Math. Soc. Colloquium Publ. 15, New York, 1932.

36

Anda mungkin juga menyukai