Cstar Algebras JAE
Cstar Algebras JAE
J. A. Erdos
Department of Mathematics
King’s College
London WC2R 2LS
ENGLAND
Abstract
In these notes we shall use the term algebra to mean a linear associative
algebra where, if no field is mentioned, the scalars will be the complex field
C. An algebra A is said to be a normed algebra if it has a norm that makes
it into a normed linear space and the norm also satisfies
If A is a normed algebra and A is a Banach space (that is, A, with its norm,
is complete) then A is called a Banach algebra.
∗
An involution on A is a map : A → A (a 7→ a∗ ) such that
(iii) a∗∗ = a,
(v) (ab)∗ = b∗ a∗ .
ab − a0 b0 = (a − a0 )(b − b0 ) + (a − a0 )b0 + a0 (b − b0 )
John Erdos 3
to obtain
a∗ a∗
Clearly sup||x||≤1 ||ax|| ≤ sup||x||≤1 ||a||.||x|| ≤ ||a|| and when x = ||a∗ ||
= ||a||
,
this supremum is attained.
4. If a C∗ -algebra A has an identity e, then e = e∗ .
In any algebra, the identity is unique. Now for all a ∈ A,
It is a simple exercise, using the theorems of Fubini and Tonelli to show that
||f ∗ g||1 ≤ ||f ||1 .||g||1 ,
John Erdos 5
(see Lemma 3.2 for a proof of a slightly more general result). It follows that
L1 (R) is a commutative Banach algebra. It is not a C∗ -algebra (although one
can define an isometric involution by f ∗ (t) = f (−t)).
L1 (T), the integrable functions on the unit circle is defined similarly. We can
parametrise the unit circle T by {eit : 0 ≤ t < 2π} and identify functions on
T with periodic functions of period 2π on R. The norm and multiplication
are defined as above, except that the limits of integration are 0 and 2π.
P∞
`1 (Z) is the space of sequences {ξn : −∞ < n < ∞} such that −∞ |ξn |
converges. The norm of a sequence x = (ξn ) is given by
∞
X
||x|| = |ξn |
−∞
All the above classical examples are special cases of the following situation.
A topological group is a group with a topology on it that makes the group
operations continuous. Let G be a topological group which is abelian as
a group and locally compact as a topological space. It is shown in books
on measure theory that on any such group one can find a measure µ that
is invariant under the group operation. That is to say, for any measurable
subset δ of G,
µ(δx) = µ(δ)
for all x ∈ G. Here the group operation is written as multiplication and δx
denotes the set {yx : y ∈ δ}. This measure is unique (up to a multiplicative
constant) and is called Haar measure.
The set L1 (G) of complex-valued functions that are integrable with respect
to Haar measure forms a Banach algebra under the norm
Z
||f || = |f (t)| dt .
G
and multiplication
Z
(f ∗ g)(t) = f (s)g(s−1 t) ds .
G
It is easy to verify that all the above examples fit into this framework. It is
these examples that form the subject of abstract harmonic analysis which is
treated in more detail in Section 3.
6 C*-algebras
Looking ahead. The main aim of these notes is to prove that all C∗ -
algebras are of the form of the examples above. Commutative C∗ -algebras
are like Examples 1 or 2 (depending on whether there is an identity) and all
C∗ -algebras are of the form of Example 3.
Much of the terminology of C∗ -algebras is inherited from operator theory
on Hilbert space. Thus, an element a of a C∗ -algebra A is said to be self-
adjoint if a∗ = a, and normal if aa∗ = a∗ a. If A has an identity e then
u ∈ A is called unitary if uu∗ = u∗ u = e.
σ(a) = {λ : λe − a is singular } .
P
Since ||xn || is convergent and hence a Cauchy series, we have that (sn ) is
a Cauchy sequence thus converges. This shows that absolutely convergent
series in a Banach space are convergent.
John Erdos 7
P∞ n
Proof. Since ||xn || ≤ ||x||n , and ||x|| < 1, it is clear that 0 x is absolutely
convergent. Denote its sum by y. Now note the identity
e = (e − x)y = y(e − x)
(i) G is open,
So ³ ´³ ´
||a−1 − b−1 || 1 − ||b − a||.||a−1 || ≤ ||a−1 ||2 .||b − a|| .
First taking ||b − a|| sufficiently small to make 1 − ||b − a||.||a−1 || > 12 , it is
clear that ||a−1 − b−1 || may be made arbitrarily small by taking ||b − a|| small
enough.
Corollary 1.3 The closure of a proper ideal is proper. Hence maximal ideals
are closed.
8 C*-algebras
Proof. We first show that σ(a) is bounded. In fact, if |λ| > ||a|| then
||λ−1 a|| < 1 and so, by 1.1 λe − a = λ(e − λ−1 a) has an inverse. Hence σ(a)
is contained in the ball centre 0, radius ||a||.
To prove that σ(a) is closed, we show that ρ(a) is open. If λ ∈ ρ(a) then
λe − a ∈ G where G, the set of invertible elements, is open (1.2). Hence, for
some ² > 0, any element b with ||b − (λe − a)|| < ² is in G. Thus if |µ − λ| < ²
then µe − a ∈ G and so µ ∈ ρ(a), showing that ρ(a) is open. (More briefly
one can say that ρ(a) is open since it is the pre-image of the open set G
under the continuous map λ 7→ λe − a.)
For the proof that σ(a) 6= ∅, first note that, if λ, µ ∈ ρ(a) then
f (µ) − f (λ)
f 0 (λ) = lim
µ→λ µ−λ
( )
(µe − a)−1 − (λe − a)−1
= lim φ
µ→λ µ−λ
n o
= lim φ −(λe − a)−1 (µe − a)−1
µ→λ
= −φ(λe − a)−2 .
Note that above (and again below), we use the continuity of inversion,
(Lemma 1.2 (ii)) several times and also the continuity of φ .
John Erdos 9
"µ ¶−1 #
1 a
lim |f (λ)| = lim φ e− = 0.
|λ|→∞ |λ|→∞ |λ| λ
Note that it is essential for the above proof that the field of scalars of our
Banach algebra is the complex field.
10 C*-algebras
(ii) if 0 ∈
/ σ(a) ½ ¾
1
σ(a−1 ) = : λ ∈ σ(a) .
λ
and if all the factors (λi e − a) had inverses, so would µe − p(a) (since the
factors commute). Therefore, at least one zero λi of q is in the spectrum of
a and p(λi ) = µ; that is, µ ∈ {p(λ) : λ ∈ σ(a)}.
(ii) This is obvious from the relation
µ ¶
1
λe − a = −aλ e − a−1 .
λ
John Erdos 11
and x = αe−a β
is self-adjoint. Hence it is sufficient to show that x + ie is
invertible for every self-adjoint x.
Suppose this is not the case so that −i ∈ σ(x). Then for any ξ,
and so, from 1.4, |ξ + 1| ≤ ||ξ + ix||. Taking ξ to be real the C∗ condition
shows that
This implies that 1 + 2ξ ≤ ||x||2 for all real ξ, which is impossible. Thus
−i ∈
/ σ(x) and so the spectrum of any self-adjoint element is real.
The following is a simple result on the spectrum of a product.
It is clear from the proof of Theorem 1.4 that ν(a) ≤ ||a||. The next theorem
establishes a formula for ν(a). It is reminiscent of the formula
1 1
= lim |an | n
r n→∞
P∞
for the radius of convergence of the scalar power series n=0 an z n .
Since φ is continuous,
³ ´ ∞
X
g(z) = φ (e − za)−1 = z n φ(an ) .
n=0
Initially we only know that this power series expansion of g is valid for |z| <
1 1
||a||
. However, since g is holomorphic on the larger disc {z : |z| < ν(a) } and the
Taylor expansion of a holomorphic function is unique, the above expansion is
1 1
valid for |z| < ν(a) . Thus, for |z| < ν(a) then limn→∞ z n φ(an ) = 0. Actually,
we only need to know that this sequence is bounded, that is, for each φ in
the dual of A, there exists a constant Kφ such that
It now follows from the Uniform Boundedness Theorem that there exists a K
1
independent of φ such that, for all n and all z with |z| < ν(a) , ||z n an || < K;
that is,
1
n n1 Kn
||a || < .
|z|
1 1
It follows that ||an || n ≤ K n ν(a) and so
1
sup||an || n ≤ ν(a) .
||a|| = ν(a) .
Proof. This is obvious either from the above result or from Theorem 1.9.
Corollary 1.13 If b ∈ B, and σB (b) has empty interior, then σB (b) = σA (b)
||π(a)||2 = νB (π(aa∗ )) .
Replacing a by aa∗ in (1) and using the above we have
||π(a)||2 = νB (π(aa∗ )) ≤ ||aa∗ || = ||a||2 .
An examination of the above proof shows that the C∗ condition was used
only for the range algebra B and the theorem is valid with a weaker condition
on A, namely that it be a Banach algebra with isometric involution. The
following consequence is clear for algebras with identities.
The result above is also true for algebras without identities. It can be proved
by using the procedure given in 1.3 below to embedd each algebra into an
algebra with identity and apply Corollary 1.15. The detailsd are omitted.
(this definition is very natural if you think of (a, λ) as a plus λ times a formal
identity). Of course, A is identified with the set {(a, 0) : a ∈ A}. Note that
A is an ideal of A1 . The element (0, 1) = e is the identity of the algebra A1 .
There are many ways to define a norm on A1 to make it into a Banach
algebra. For example
k(a, λ)k = kak + |λ|
is complete and it is easy to prove that it satisfies the multiplicative norm
condition (i) for a Banach algebra. However, this does not make A1 into a
C∗ -algebra.
To make A1 into a C∗ -algebra, define a norm |||.||| as follows :
where x varies over A. Note that multiplication by (a, λ) acts as linear oper-
ator on the ideal A of A1 and |||(a, λ)||| is simply the norm of this operator.
From this it follows easily that |||.||| satisifies the multiplicative condition for
the norm of a Banach algebra.
Recall (Section 1.1, consequence 3) that kak = supkxk≤1 kaxk showing that
|||(a, 0)||| = kak, so that the embedding of A into A1 is isometric. To verify
the C∗ condition,
2 Commutative theory.
In this section we shall show that a commutative Banach algebra with iden-
tity can be mapped into the algebra of all continuous functions on some
compact space. In the case of a C∗ -algebra the embedding is an isometric
*-isomorphism.
Before embarking on the proofs, we need some remarks on quotients. Recall
that if X is a linear space and Y is a subspace of X then we define an
equivalence relation ∼ on X by
x1 ∼ x2 ⇐⇒ x1 − x2 ∈ Y .
k[x]k = inf kx + yk
y∈Y
makes X /Y into a Banach space. All the above is standard Banach space
theory.
If A is an algebra and I ⊆ A then it can be verified that
[a].[b] = [a.b]
Proof. From Corollary 1.3, M is closed. Also, A/M has no proper ideals
since if I were such then Z = ∪{a : [a] ∈ I} would be a proper ideal of A
containing M .
John Erdos 19
â(φ) = φ(a) .
Proof. This is a routine verification using the fact that each φ is a homo-
morphism. For example, to show that ab c = â.b̂, simply note that
c
(â.b̂)(φ) = φ(a).φ(b) = φ(ab) = ab(φ) .
Theorem 2.4
(and it is a fact that the two intersections are equal). When A is commutative
then,
R = ∩{all maximal two-sided ideals of A } .
Now,
â = 0 ⇔ â(φ) = 0 ∀φ ∈ Φ ⇔ a ∈ M ∀M ∈ M ⇔ a ∈ R.
that is
{ψ ∈ Φ : |âi (φ) − âi (ψ)| < ², i = 1, 2, · · · , n}
as the quantities a1 , a2 , · · · , an in A and ² > 0 range over all choices.
Convergence of sequences in this topology is given by
φn → φ ⇔ â(φn ) → â(φ) in C ∀a ∈ A
⇔ φn (a) → φ(a) ∀a ∈ A
22 C*-algebras
Lemma 2.5 The carrier space of a commutative Banach algebra with iden-
tity is compact.
Proof. Let Φ be the carrier space of A and let φ ∈ Φ. From Theorem 2.4 (iii),
φ(a) ∈ σ(a) and so |φ(a)| ≤ ν(a) ≤ kak. Thus the closed disc Za in C, centre
0 radius kak is compact and contains φ(a) for all a ∈ A. Let
Y
Z= Za .
a∈A
The elements of Φ are those elements of Z that also satisfy the algebraic
properties of homomorphisms. Thus if φ ∈ Φ then for all a, b ∈ A,
Then, directly from the definition of the product topology it is clear that La,b
is continuous and hence its kernel L−1a,b (0) is closed. The algebraic condition
on φ given above is that φ is in the kernel of each La,b . Similarly we define,
for a, b ∈ A and λ ∈ C,
and all these functions are continuous. It is easy to see that an element of Z
is in Φ exactly when it is in the intersection of the kernels of all the functions
La,b , Ma,b , Sa,λ , I as a, b, λ take all possible values. Therefore it is a closed
subset of Z and so it is compact.
The theory developed in the results above, when put together, constitute a
proof of the following important theorem.
Lemma 2.7 If A is generated by the identity e and one element a then the
carrier space Φ of A is homeomorphic to σ(a).
from Theorem 2.4 (iii) it is onto σ(a). To show that â is injective we use our
hypothesis that the elements {p(a)}, as p ranges over all polynomials, form
a dense subset of A. Then
â(φ1 ) = â(φ2 ) ⇒ φ1 (a) = φ2 (a)
⇒ φ1 (p(a)) = φ2 (p(a)) ∀ polynomials, since φ1 , φ2 are homomorphisms
⇒ φ1 (x) = φ2 (x) ∀x ∈ A, since φ1 , φ2 are continuous
⇒ φ1 = φ2
Proof. Let Φ be the carrier space of A and let a 7→ â be the Gelfand map
as above. Clearly every element of a commutative C∗ -algebra is normal and
so from Corollary 1.10 the norm of each element of A is equal to its spectral
radius. Thus kak = kâk.
Now A is a Banach space and thus complete. Therefore, since the Gelfand
map is isometric, {â : a ∈ A} is a complete subset of C(Φ) and so closed.
We shall use the Stone-Weierstrass Theorem to show that it is the whole of
C(Φ).
We show that the Gelfand map preserves the involution (where the involution
in C(Φ) is defined in the natural way : f ∗ (φ) = f (φ). For a ∈ A we write
a = x + iy where x = 12 (a + a∗ ), y = 2i1 (a − a∗ ) are selfadjoint. From
Theorem 1.7 (iii), x and y have real spectra and so, since ˆ(φ) ∈ σ(x), the
functions x̂ and ŷ are real-valued. Therefore
This shows that the image of  is a *-subalgebra and also that the Gelfand
map preserves the adjoint operation. The remaining conditions needed to
apply the Stone-Weierstrass Theorem have already been proved, namely
that  separates the points of Φ and contains the unit function (Theo-
rem 2.4 (i), (v)). Therefore  = C(Φ).
â(φ1 ) = â(φ2 ) ⇒ ac∗ (φ1 ) = (â(φ1 ))∗ = â(φ1 ) = â(φ2 ) = ac∗ (φ2 )
φ1 (p(a, a∗ )) = φ2 (p(a, a∗ ))
for all polynomials in two variables. Thus, by the hypothesis, φ1 (x) = φ2 (x)
for all x ∈ A so that φ1 = φ2 . The remainder of the proof is exactly the same
as that of Lemma 2.7.
26 C*-algebras
The observations following Lemma 2.7 also apply to Lemma 2.9 above. In
view of Theorem 2.8 the conclusion here is that A is isometrically isomorphic
to C(σ). Once again, the image of the generator a is the identity function
f (s) = s
Φ = Φ1 \ {φ∞ }
Since the identity of A1 is (0, 1), it follows that φ∞ ((a, λ)) = λ. Also, ev-
ery element of Φ is (by restriction) a non-zero homomorphism of A into C.
Conversely, if ψ is a non-zero homomorphism of A into C, then φ : A1 → C
defined by φ((a, λ)) = ψ(a) + λ is easily shown to be an element of Φ1 . Thus
Φ is, as before, the space of all non-zero homomorphisms of A into C and is
called the carrier space of A. Since Φ1 is compact, the topology induced on
Φ is locally compact (i.e. every point has a compact neighbourhood).
From Theorem 2.8, the Gelfand map (a, λ) 7→ (a, dλ) is an isometric iso-
morphism of A1 onto C(Φ1 ). The composition of this with the embedding
A → (A, 0) takes A onto the algebra of all functions of C(Φ1 ) that vanish at
φ∞ . Restricting the functions to Φ shows that A is mapped onto an algebra
 of continuous functions on Φ.
A function f ∈ C(Φ1 ) satisfies f (φ∞ ) = 0 if and only if, given any ² > 0,
there exists an open neighbourhood N of φ∞ such that |f (φ)| < ² for φ ∈ N .
Since the complement of N is compact, this can be rephrased in terms of
elements of Φ alone as follows : given any ² > 0, there exists a compact set
K such that |f (φ)| < ² for all φ ∈/ K. This is exactly the definition of a
“function vanishing at infinity” on a locally compact space.
John Erdos 27
Let H be a Hilbert space and let B(H) be the C∗ -algebra of all bounded
linear operators on H. It T ∈ B(H) is a normal operator then T T ∗ = T ∗ T
and so the algebra A generated by T, T ∗ and I (the identity operator) is a
commutative C∗ -subalgebra of B(H).
Let σ be the spectrum of T in B(H). Then Theorem 1.16 shows that
σA (T ) = σ .
It follows from Lemma 2.9 that the carrier space of A is homeomorphic to
σ. Using the remarks following Lemma 2.7 and Lemma 2.9 we see that
the Gelfand map composed with this homeomorphism gives an isometric *-
isomorphism of A onto C(σ). Moreover, the homeomorphism can be chosen
so that T is mapped onto the identity function; i.e. onto the function fT
where fT (t) = t. If A ∈ A, we denote the image of A under this map by fA
and if f ∈ C(σ), we denote the operator mapped onto f by Af .
The spectral theorem for T is now easily derived using two standard the-
orems. Each of these is usually referred to as the “Riesz representation
theorem”. The first is the description of the dual of C(σ) as the set of Borel
measures on σ and the second is the correspondence between sesqui-linear
functionals on H × H and bounded operators on H.
First we note that, for any x, y ∈ H, the map φ : C(σ) → C defined by
φ(f ) = hAf x, yi
is a continuous linear functional. Thus there exists a Borel measure µx,y on
σ such that, Z
hAf x, yi = f (t) dµx,y .
σ
We shall frequently use the fact that if two measures give the same integral
of continuous functions then they are the same. If f ∈ C(σ), and |f (t)| ≤ 1
then kAf k = kf k ≤ 1 and so
¯Z ¯
¯ ¯
¯ f (t) dµx,y ¯ = |hAf x, yi| ≤ kxk.kyk
¯ ¯
σ
which, by a standard argument using regularity, shows that for all Borel sets
δ,
|µx,y (δ)| ≤ kxk.kyk.
28 C*-algebras
Also, it is easy to verify that for a fixed Borel set δ, µx,y (δ) is linear in x and
conjugate linear in y. For example, if f ∈ C(σ)
Z
f dµx+x0 ,y = hAf (x + x0 ), yi = hAf x, yi + hAf x0 , yi
Z
= f d(µx,y + µx0 ,y )
showing that µx+x0 ,y (δ) = µx,y (δ) + µx0 ,y (δ) for any Borel set δ.
The next step is to extend the definition of Af to all functions f in the set
B(σ) of bounded Borel measurable functions on σ. To do this, note that for
f ∈ B(σ), Z
φ(x, y) = f dµx,y
σ
We now show that the extended map f 7→ Af preserves the algebraic struc-
ture, that is, it is a *-homomorphism.
Since for f ∈ C(σ) we have that A∗f = Af¯, it follows easily that
Z Z Z
f dµx,y = hAf x, yi = hA∗f y, xi = f¯ dµy,x = f dµy,x .
σ σ σ
This shows that integrating with respect to µx,y is the same as integrating
with respect to µy,x . The relation therefore holds for all f ∈ B(σ). Thus
µx,y = µy,x and consequently A∗f = Af¯ for all f ∈ B(σ).
The proof that f 7→ Af preserves multiplication requires two stages. If
f, g ∈ C(σ), Af g = Af .Ag and so, for all x, y
Z Z
f g dµx,y = f dµAg x,y .
σ σ
So “g dµx,y = dµAg x,y ” and the above relation holds for all f ∈ B(σ). Thus
for all f ∈ B(σ) and g ∈ C(σ),
Af g = Af .Ag .
John Erdos 29
Also
Af g = (Af¯ḡ )∗ = (Af¯Aḡ )∗ (Aḡ )∗ (Af¯)∗ = Ag Af .
But then
Z Z
gf dµx,y = hAgf x, yi = hAg Af x, yi = g dµAf x,y
σ σ
and the equality of the integrals for all g ∈ B(σ) follows as before. Thus for
all f, g ∈ B(σ) we have Af Ag = Af g = Ag Af .
The isometry of the Gelfand map is not extended to f ∈ B(σ), since f may
be large on sets of measure 0. However, it is easy to verify that kAf k ≤ kf k
for f ∈ B(σ).
We now look at the operators Af for the cases when f is the characteristic
function χδ of some Borel set δ. We write E(δ) = Aχδ . Clearly each E(δ)
commutes with T and E(σ) = I. We show that E(·) is an instance of what is
called a “spectral measure”. Since (χδ )2 = χδ and χδ is real, it is immediate
that E(δ)2 = E(δ) and E(δ)∗ = E(δ) so that each E(δ) is an orthogonal
projection on H. It is easy to see that the following properties hold :
1. E(α ∩ β) = E(α).E(β)
We extend the definition of E(δ) to all Borel subsets of C by E(δ) = E(δ ∩σ).
Clearly all the properties are preserved. A map from the Borel subsets of
C with values in the orthogonal projections on Hilbert space H satisfying
30 C*-algebras
Z n
X
f (λ)E(dλ) = αi E(δi ) .
σ i=1
R
Clearly for a simple function f , k σ f (λ)E(dλ)k ≤ kf k. Now for every
f ∈ B(σ), there is a sequence (fk ) of simple functions
R
converging uniformly
to f on σ. It follows easily that the sequence ( σ fk (λ)E(dλ)) converges in
the norm of B(H) to an operator which is dependent only on f and not the
choice of the sequence (fk ). Define
Z Z
f (λ)E(dλ) = lim fk (λ)E(dλ) .
σ k→∞ σ
R
We clearly would like to have that σ f (λ)E(dλ) = Af and this is indeed the
case. To prove it we first note that the equality is true for simple functions.
Now if (fk ) is a sequence of simple functions converging uniformly to f , since
kAf − Afk k ≤ kf − fk k it is clear that Af equals the integral. Looking at the
special case when f is the identity function, we obtain the spectral theorem.
Spectral Theorem for Normal Operators. For a normal operator on
a Hilbert space H, there exists a spectral measure E(·) with support the
spectrum of T such that Z
T = λ E(dλ) .
We shall not go further with this topic in any detail in these notes. However
we should mention a few developments.
R
1. It follows that defining f (T ) = σ f (λ)E(dλ) gives a functional calculus
for bounded Borel functions of a normal operator. In fact one could
John Erdos 31
deal with functions that are essentially bounded (in the appropriate
sense) but we will not go into these technicalities.
3. One can prove that, for any Borel set δ, the spectrum of T E(δ), re-
garded as an operator on the range of E(δ), is δ. The converse to
this is true and shows the uniqueness of the spectral measure of an
operator: namely if E(·) is a spectral measure commuting with T such
that σ(T E(δ)) ⊆ δ for all Borel subsets δ of C, the E(·) is the spectral
measure of T .
The purpose of this section is to connect the Gelfand theory and Fourier
analysis on groups. The most familiar and the most important groups for
Fourier analysis are the real line R, euclidean n-space Rn , the integers Z (all
with addition as the operation) and the circle group T of complex numbers
of modulus 1 under multiplication (equivalently (0, 2π] with addition modulo
2π). We shall refer to these groups as “our specific groups”.
The appropriate setting for elementary abstract harmonic analysis is to con-
sider functions on a locally compact abelian group G (the more advanced
portions of the subject deal with more general groups). We have seen that
such a group G has a Haar measure and that L1 (G), with convolution as
multiplication, is a Banach algebra. We shall introduce the statements of
the results in this generality. However, when the proofs are much easier (or
even trivial) in the case of our specific groups mentioned above, we shall omit
the general proofs.
For the whole of this section, G will denote a locally compact abelian group
and all integrations will be with respect to Haar measure over the whole
group. Recall (Section 1, Example 5) that L1 (G) is a Banach algebra with
multiplication Z
(f ∗ g)(t) = f (s)g(s−1 t) ds .
G
Proof. For the general case one needs the fact that given any subset δ of G,
then δ and δ −1 = {x−1 : x ∈ δ} have the same Haar measure. For our specific
groups this fact is clear and the rest of the proof is a trivial verification using
a change of variable in the integration.
kf ∗ gkp ≤ kf k1 kgkp .
Proof. The proof uses the theorems of Fubini and Tonelli on double inte-
p
gration. Let q = p−1 be conjugate to p and let h be an arbitrary element
of L (G). Then f (x)g(x−1 y)h(y) is measurable on G × G. So, using the
q
Z Z Z ½Z ¾ 1 ½Z ¾1
p q
−1 −1 p q
|f (x)g(x y)h(y)| dy dx ≤ |f (x)| |g(x y)| dy |h(y)| dy dx
= kf k1 kgkp khkq
kf ∗ gkp ≤ kf k1 kgkp .
but we shall not formally introduce this. We indicate proofs for our specific
groups.
In `1 (Z), the sequence e = (²n ) where ²0 = 1 and ²n = 0 for n 6= 0 acts as an
identity (and also e ∈ `2 (Z)). Thus if x = (ξn ) 6= 0, Tx e = x 6= 0.
The proof for L1 (R) is similar to the general case and involves some tech-
nicalities. We first deal with a continuous function f of compact support.
If f 6= 0, choose ² > 0 such that ² < kf k1 . Clearly it is enough to find a
function g ∈ L2 (R) ∩ L1 (R) such that kf ∗ g − f k1 < ². Since f vanishes
outside a compact set, there exists k > 0 such that f (x) = 0 for |x| > k.
Also an elementary result shows that f is uniformly continous so there is a
²
δ > 0 with δ < 1 such that for |s| < δ and all t, |f (t − s) − f (t)| < 2(k+1) .
The condition δ < 1 is imposed so that if |s| < δ, |f (t − s) − f (t)| vanishes
for |t| > k + 1. Therefore, for |s| < δ,
Z
|f (t − s) − f (t)| dt < ² .
Define g by (
1
if |s| < δ
g(t) = 2δ
0 otherwise.
R R
Then g(s) ds = 1 so f (t) = f (t)g(s) ds and
Z ¯Z ¯
¯ ¯
kf ∗ g − f k1 = ¯ [f (t − s) − f (t)]g(s) ds¯ dt
¯ ¯
Z Z
≤ |f (t − s) − f (t)| dt g(s) ds
< ².
For an arbitrary function f ∈ L1 (R), we use the fact that the continuous
functions of compact support are dense in L1 (R). It then follows by a routine
approximation argument that, for f 6= 0, the operator Tf is non-zero.
The proof for L1 (T) (and indeed for a general locally compact abelian group
G) follows along the same lines. Instead of the interval [−δ, δ] one needs a
suitable compact neighbourhood of 1. The details are omitted.
and so φξ is a homomorphism.
Since ξ is a non-zero function, clearly φξ 6= 0. Similarly, it is clear that the
map ξ 7→ φξ is an injection.
To show that ξ 7→ φξ is a surjection, we use the fact that the dual of L1 is
L∞ . Suppose that φ ∈ Φ. Then φ is a continuous linear functional on L1 (G)
and so, for some α ∈ L∞ (G),
Z
φ(f ) = f (x)α(x) dx .
Thus Z
φ(gy−1 )
φ(f ) = f (y) dy.
φ(g)
φ(gy−1 )
ξ(y) = ,
φ(g)
φ(g −1 )
Note that it is implicit in the above proof that φ(g) y
is independent of of
the choice of g (provided φ(g) 6= 0). This can also be seen from the relation
f ∗ gy = fy ∗ g. Note also that ξ(y) = ξ(y −1 ) = φ(g
φ(g)
y)
.
The group Ĝ inherits a topology from the Gelfand topology on Φ via the
bijection of the above theorem. That is, a set δ is open in Ĝ when {φξ : ξ ∈ δ}
is open in Φ. It is a fact that this makes Ĝ into a topoplogical group (i.e.
with this topology, the group operations are continuous). We shall verify this
for our specific groups but the general proof will not be given in these notes.
We call the group Ĝ with this topology the dual group of the group G.
For f ∈ L1 (G), the Gelfand transform of f is a function with domain Φ, the
carrier space of L1 (G). Since Φ and Ĝ are homeomorphic, we may therefore
regard the transform of f as a function on Ĝ. Thus
fˆ(ξ) = φξ (f )
∗ g = fˆ.ĝ
1. fd
3. kfˆk∞ ≤ kf k1 .
We now examine our specific groups and will show that the above indeed
gives the classical Fourier transform. The results quoted should be familiar
in the case of the classical theory.
The additive group R of real numbers.
We first identify the characters. If ξ is a character, |ξ(x)| = 1 for all x ∈ R,
ξ(0) = 1 and ξ is continuous. Therefore there exists a δ such that | arg ξ(x)| <
π/2 for |x| < δ.
John Erdos 37
Since dyadic rationals are dense in R, using the continuity of ξ we have that
the above hold for all real numbers α. Now let y = θt . The for any x ∈ R
ξ(x) = eixy
for some y ∈ R.
Conversely it is easy to see that distinct y ∈ R give distinct characters. Also
it is clear that multiplication of characters corresponds to addition of the
numbers that give rise to them. Thus, as far as algebraic properties are
concerned, the dual group of R is R.
We now verify that the topology induced on R by the Gelfand topology
coincides with the usual topology. Note that a base for the neighbourhoods
of 0, (that is, of the character ξ0 corresponding to the map x 7→ eixy when
y = 0, hence ξ0 (x) ≡ 1) in the Gelfand-induced topology is given by
½ ¯Z ¯ ¾
¯ ixy
¯
N (ξ0 , f1 , f2 , · · · , fn ; ²) = ¯ ¯
y : ¯ fr (x)(1 − e ) dx¯ < ², 1 ≤ r ≤ n
R
for all choices of f1 , f2 , · · · , fn and ². Since the map y →
7 f (x)(1 − eixy ) dx
is continuous, it is clear that every neighbourhood of this type contains an
38 C*-algebras
1 Z∞
fˆ(y) = √ f (x) e−ixy dy .
2π −∞
The properties 1 – 3 above are then familiar results of Fourier theory, Prop-
erty 2 being the Riemann-Lebesgue lemma.
The additive group of Rn .
Note that each subspace of Rn is a subgroup. If ξ is a character of Rn , it
induces a character on each one-dimensional subspace. Let {er : 1 ≤ r ≤ n}
be the usual basis of Rn . Then x 7→ ξ(xer ) is a character of R and so from
above we have that for some yr ,
ξ(xer ) = eixyr .
Thus, if x = (x1 , x2 , · · · , xn ),
³X ´
ξ(x) = ξ xr er = eix.y
where y ∈ Rn and x.y is the usual inner product. It follows easily that the
dual of Rn is Rn and the Gelfand transform of f ∈ L1 (Rn ) is therefore
à !n Z
1
fˆ(y) = √ f (x)e−ix.y dx ,
2π n
R
ξ(x) = eixy .
John Erdos 39
The doubly infinite sequence {fˆ(n)} is usually called the sequence of Fourier
coefficients of f .
The group Z of integers.
Once again, the characters are of the form ξ(n) = einy but now any two
numbers differing by 2π give the same character. Thus the dual of Z is T.
(The proof that the topology of T arising from the Gelfand topology is the
usual topology is an easy verification).
Haar measure on Z gives unit mass to every point and so L1 (Z) is the space
of absolutely convergent doubly infinite sequences. The Fourier transform of
a member (an )−∞<n<∞ of L1 (Z) is the function f ∈ C(T) given by,
∞
X
f (y) = an einy
−∞
1
In a more general context this would be called a *-representation of A on a Hilbert
space.
John Erdos 41
Note that in many other treatments ρ(e) = I is not included in the definition
of the term representation. However, the condition is convenient and, as will
be seen later, there is no essential loss in generality.
A representation is said to be faithful if it is an injection. Note that Theo-
rem 1.14 shows that (with the above definition) any representation is auto-
matically contiuous with kρ(a)k ≤ kak. The following lemma is included to
motivate the development.
f (a) = hρ(a)ξ, ξi
satisfies
(iii) f (e) = 1,
(iv) f is continuous
(v) kf k = 1 .
Proof. The only part that is not completely trivial is (ii) whose proof is as
follows :
Parts (iv) and (v) follow from the fact (Theorem 1.14 ) that kρ(a)k ≤ kak.
Proof. These results arise from the fact that φ(x, y) = f (y ∗ x) is a non-
negative sesquilinear form on A. The details follow.
(i) This is proved using polarization. We have
Choosing λ = ke−iθ where θ = arg f (b∗ a) with k real gives, using (i), that
for all k ∈ R, and the result merely states the fact that the discriminant of
the above quadratic is negative.
Corollary 4.4
Proof. Suppose a = a∗ and kak ≤ 1. Then σ(a) ⊆ [−1, 1] and the spectral
mapping theorem shows that σ(e−a) ⊆ [0, 2]. Thus by Lemma 4.1 A contains
a selfadjoint element b such that b2 = e − a. Therefore f (e − a) = f (b∗ b) ≥ 0
and so
f (a) ≤ f (e) .
x∗ x
Now if x ∈ A, applying the above to and using Corollary 4.4 (ii) we
kx∗ xk
have that
f (e) = |f (e)| ≤ kf k
and so kf k = f (e).
Note that the only time that the C∗ condition, via Lemma 4.1, is used in this
section is in the above proof. By justifying the formal binomial expasion of
1
(e − x) 2 for kxk ≤ 1, the above lemma can be proved for Banach *-algebras
with isometric involution. This shows that the GNS construction is valid for
these more general algebras.
The following lemma gives an inner product space on which the GNS con-
struction will provide a representation.
We now show that f (b∗ a) depends only on the equivalence classes of a and
b. For all x, y ∈ N , using Lemma 4.3 (i) and the above,
|f (y ∗ xy)| ≤ kxkf (y ∗ y) .
which proves that ρf (a)∗ = ρf (a∗ ) on the dense subspace A/N of Hf and
hence, since both ρf (a)∗ and ρf (a∗ ) are bounded, they are equal.
We have thus proved that ρf is a representation of A on Hf . Also
hρf (a)[e], [e]i = h[a], [e]i = f (a)
and, since f is a state, k[e]k2Hf = f (e∗ e) = f (e) = 1 so [e] is a unit vector.
This completes the proof.
the study of elements of the form a∗ a. The major tool in the investigation
of this is the commutative Gelfand-Naimark theorem.
An element a of A is said to be positive if it is selfadjoint and its spectrum
consists of non-negative real numbers. The main technical problem will be
to show that every element of the form x∗ x is positive. Note that, from
Theorem 1.16 , if B is a C∗ -subalgebra of A, with the same identity as A,
then σB (b) = σA (b) for all b ∈ B. Thus an element a is positive in A if
and only if it is positive in any one *-subalgebra containing it and so it is
sufficient to look at its spectrum in the C∗ -algebra generated by a and e.
The set of all positive elements of A will be denoted by A+ . Note that the
spectral mapping theorem (Theorem 1.6 ) shows that if a ∈ A+ then an ∈ A+
for all positive integers n.
John Erdos 47
(i) ke − ak ≤ 1 ⇒ a ∈ A+ ,
(ii) a ∈ A+ , kak ≤ 1 ⇒ ke − ak ≤ 1 ,
° °
(iii) a ∈ A+ ⇐⇒ °°kake − a°° ≤ kak .
λa + µb
and so by part (i) of Lemma 4.9 , ∈ A+ . Consequently, λa + µb ∈
kak + kbk
A+ and so A+ is a convex cone.
If x ∈ A+ ∩ (−A+ ) then x is selfadjoint and σ(x) = {0}. Hence ν(x) = 0 and
so from Corollary 1.10, x = 0.
To show that A+ is closed, note that the set S of all selfadjoint elements of
A is closed. Lemma implies that A+ is a closed subset of S and so A+
is closed.
48 C*-algebras
(i) a ∈ A+ ,
Proof. Lemma 4.1 establishes (i)⇒ (ii) and (ii) ⇒ (iii) is trivial. The
substance of the Theorem is to prove that (iii) ⇒ (i).
If a = y ∗ y then a is self-adjoint. Let Φ be the carrier space of the commutative
C∗ -algebra C generated by a and e. Then â is a real function on Φ and,
writing f (φ) = max[â(φ), 0] and g(φ) = − min[â(φ), 0], we have that f and
g are non-negative functions on Φ with
â = f − g and f g = gf = 0 .
a = y∗y = b − c
Also
cy ∗ (cy ∗ )∗ = cy ∗ yc = cbc − c3 = −c3 ∈ −A+ . (2)
Therefore, using (1), (2) and Theorem 4.10 ,
However, since (cy ∗ )∗ cy ∗ and cy ∗ (cy ∗ )∗ have the same non-zero numbers in
their spectrum (Lemma 1.8), (3) implies that
cy ∗ (cy ∗ )∗ = −c3 ∈ A+ .
John Erdos 49
Proof. We show that f is positive and therefore a state. First we show that
if a is selfadjoint then f (a) is real. We may suppose that kak ≤ 1. Then, for
any real k, using the C∗ - condition,
Proof. We first prove the theorem for the case of a C∗ -algebra A with identity
e. For each non-zero x ∈ A let fx be a state (as found in the preceding lemma)
such that fx (x∗ x) 6= 0. Let ρx be the representation of A on a Hilbert space
Hx constructed from fx by the GNS construction Theorem 4.8 . Then
and so ρx (x) 6= 0.
Let H = ⊕{Hx : x ∈ A} and define ρ : A → B(H) by