an invitation to c -algebras* - strung mediations and ...strung.me/karen/the_fringe/invitation to...

67
AN INVITATION TO C * -ALGEBRAS* KAREN STRUNG 0. Background and context The study of C * -algebras fits into the broader mathematical framework of op- erator algebras. A phrase one often hears in relation to operator algebras is that they are the “noncommutative” version of classical objects like topological spaces, measure spaces, groups, and so forth. The study of operator algebras started at the beginning of the 20th century. Of course, some noncommutative mathematics, in particular the study of matrices as well as the perhaps more abstract linear operators on finite-dimensional vector spaces, was already known and reasonably developed. Indeed, matrix algebras are themselves (important) examples of oper- ator algebras. However, it is reasonable to say that the subject was brought into being by von Neumann’s efforts to make mathematically sound the emerging area of quantum physics. The noncommutative nature that popped up in places such as Heisenberg’s uncertainty principal, led von Neumann to define an abstract Hilbert space (particular Hilbert spaces, though they were not called that, were already known, in particular L 2 and 2 were studied by various people) and undertake a study of operators on such spaces. In particular, he introduced self-adjoint sub- algebras of bounded operators that were closed in the weak * -topology; these later became known as von Neumann algebras (the term was coined by Dixmier) and sometimes still as W * -algebras. Subsequent papers by Murray and von Neumann further developed the subject. C * -algebras, which are the topic of this course, have their origins in the studies of Gelfand and Naimark in the 1940’s. While it turns out that von Neumann algebras are a particular type of C * -algebra, they are often studied separately; indeed, there will not be much to say about von Neumann algebras in these notes. Gelfand and Naimark showed that, given a compact Hausdorff space X , the algebra of continuous functions on that space, C (X ), can be given an involution as well as norm which makes it a Banach * -algebra. Moreover, this norm satisfies the C * -equality: kf * f k = kf k 2 for every f C (X ). The C * -equality is precisely what is needed to pass from a Banach * -algebra to a C * -algebra. This apparently minor requirement has wide-reaching implications for the structure of C * -algebras; we shall see just what this means in the sequel. Further work by Gelfand and *or rather, Another Invitation to C * -algebras, for those who either missed or turned down Arveson’s 1998 invitation. 1

Upload: dinhhanh

Post on 16-Mar-2018

218 views

Category:

Documents


2 download

TRANSCRIPT

AN INVITATION TO C∗-ALGEBRAS*

KAREN STRUNG

0. Background and context

The study of C∗-algebras fits into the broader mathematical framework of op-erator algebras. A phrase one often hears in relation to operator algebras is thatthey are the “noncommutative” version of classical objects like topological spaces,measure spaces, groups, and so forth. The study of operator algebras started atthe beginning of the 20th century. Of course, some noncommutative mathematics,in particular the study of matrices as well as the perhaps more abstract linearoperators on finite-dimensional vector spaces, was already known and reasonablydeveloped. Indeed, matrix algebras are themselves (important) examples of oper-ator algebras. However, it is reasonable to say that the subject was brought intobeing by von Neumann’s efforts to make mathematically sound the emerging areaof quantum physics. The noncommutative nature that popped up in places such asHeisenberg’s uncertainty principal, led von Neumann to define an abstract Hilbertspace (particular Hilbert spaces, though they were not called that, were alreadyknown, in particular L2 and `2 were studied by various people) and undertake astudy of operators on such spaces. In particular, he introduced self-adjoint sub-algebras of bounded operators that were closed in the weak∗-topology; these laterbecame known as von Neumann algebras (the term was coined by Dixmier) andsometimes still as W ∗-algebras. Subsequent papers by Murray and von Neumannfurther developed the subject.

C∗-algebras, which are the topic of this course, have their origins in the studiesof Gelfand and Naimark in the 1940’s. While it turns out that von Neumannalgebras are a particular type of C∗-algebra, they are often studied separately;indeed, there will not be much to say about von Neumann algebras in these notes.Gelfand and Naimark showed that, given a compact Hausdorff spaceX, the algebraof continuous functions on that space, C(X), can be given an involution as wellas norm which makes it a Banach ∗-algebra. Moreover, this norm satisfies theC∗-equality: ‖f ∗f‖ = ‖f‖2 for every f ∈ C(X). The C∗-equality is preciselywhat is needed to pass from a Banach ∗-algebra to a C∗-algebra. This apparentlyminor requirement has wide-reaching implications for the structure of C∗-algebras;we shall see just what this means in the sequel. Further work by Gelfand and

*or rather, Another Invitation to C∗-algebras, for those who either missed or turneddown Arveson’s 1998 invitation.

1

STRUNG – INVITATION TO C∗-ALGEBRAS 2

Naimark, together with Segal, established way of constructing a representationof any C∗-algebra as a norm-closed self-adjoint subalgebra of bounded operatorson a Hilbert space. Gelfand and Naimark showed that in fact there is always afaithful representation; thus every C∗-algebra is isomorphic to such a subalgebraof operators.

Since those early days, the study of C∗-algebras has taken off in many directions.It continues to find utility in the study of physics (including quantum gravity, quan-tum information, statistical mechanics), the study of topological groups as well asthe development of quantum groups, dynamical systems, K-theory, and of course,noncommutative geometry. Some of the state-of-the-art will have been introducedat the master classes. Here, we concentrate on developing the theory from thebeginning, with minimal background requirements save for some knowledge offunctional analysis and Hilbert spaces.

STRUNG – INVITATION TO C∗-ALGEBRAS 3

Contents

0. Background and context 1

1. Banach algebras and spectral theory 4

Spectrum 4

Exercises 9

2. The Gelfand representation 11

Exercises 14

3. C∗-algebra basics 14

Minimal unitisation 16

Exercises 20

4. Positive elements 21

Approximate units and hereditary C∗-subalgebras 25

Exercises 28

5. Positive linear functionals and representations 30

Exercises 39

6. Further examples of C∗-algebras 39

UHF algebras and AF algebras 39

Group C∗-algebras 45

Crossed products 49

Universal C∗-algebras 51

Exercises 53

7. A very short introduction to classification for simple nuclear C∗-algebras 55

8. Extra material 61

Multiplier algebras 61

More about representations 62

Tensor products for C∗-algebras 63

Index 65

References 67

STRUNG – INVITATION TO C∗-ALGEBRAS 4

1. Banach algebras and spectral theory

A Banach algebra is an algebra A, together with a submultiplicative norm ‖ · ‖ :A→ [0,∞), which is complete with respect to the norm. For the purposes of thiscourse, we will consider only algebras over C.

Note that A is not necessarily unital. If A has a unit 1A then we require that‖1A‖ = 1 and in this case we call A a unital Banach algebra.

If B ⊂ A is a subalgebra, then its closure with respect to the norm of A is alsoa Banach algebra.

All Banach we will consider will be C-algebras, though they can also be definedover other fields.

1.1 Examples: (a) Let X be a topological space and let

Cb(X) := f : X → C | f continuous, bounded .

Then Cb(X) is a Banach algebra when equipped with pointwise operations andsupremum norm

‖f‖∞ = sup|f(x)| | x ∈ X.

(b) Let X be a Banach space. Let L(X) = T : X → X | T linear, continuous equipped with pointwise addition and composition for multiplication. L(X) is aBanach algebra with the operator norm

‖T‖ = supx∈X,‖x‖≤1

‖T (x)‖.

(c) Let (X,Σ, µ) be a measure space. Let

L∞(X,Σ, µ) := f : X → C | f measurable, and ∃K > 0 s.t. µ(x | |f(x)| > K) = 0.

Define a norm on L∞(X,Σ, µ) by

‖f‖ = inff=g a.e.µ

supx∈X|g(x)|.

Then L∞(X,Σ, µ) is a Banach algebra.

Spectrum. Let p(z) = λ0 +λ1z+ . . . λnzn, λi ∈ C, be a polynomial in the algebra

of polynomials in one indeterminate, C[z]. Let A be a unital algebra and a ∈ Aand denote by p(a) the element λ01A + λ1a+ . . . λna

n.

Let A be a unital algebra. An element a ∈ A is invertible if there is a b ∈ Asuch that ab = ba = 1A. In this case we write b = a−1. (This makes sense becauseif such a b exists, it is is unique. Why?)

1.2 Exercise: Let A be a unital algebra and show that the set of invertibleelements Inv(A) := a ∈ A | a is invertible in A is a group under multiplication.

STRUNG – INVITATION TO C∗-ALGEBRAS 5

The spectrum of an element a in the unital algebra A is defined to be

sp(a) = λ ∈ C | λ · 1A − a /∈ Inv(A).

1.3 Exercise: What is sp(a) for a ∈Mn and sp(f) for f ∈ C(X), where X is acompact Hausdorff space?

1.4 Suppose that 1−ab is invertible with inverse c. Then one can check that 1−bais also invertible with inverse given by 1 + bca. As a result, we have that, for anya, b in a unital Banach algebra A,

sp(ab) \ 0 = sp(ba) \ 0.

We have the following spectral mapping property for polynomials.

1.5 Theorem: Let A be a unital algebra, a ∈ A and p a polynomial in C[z].Suppose that sp(a) 6= ∅. Then sp(p(a)) = p(sp(a)).

Proof: If p is constant, the result is obvious, so we may assume otherwise. Letµ ∈ C and consider the polynomial p − µ. Since every polynomial over C splits,we can write

p(z)− µ = λ0(λ1 − z) · · · (λn − z)

for some n ∈ N \ 0, λ0, . . . , λn ∈ C and λ0 6= 0. If µ /∈ sp(p(a)) then p(a)− µ isinvertible and hence each λi − a, i = 1, . . . , n, is invertible. Conversely, it is clearthat if each λi − a is invertible, so is p(a) − µ. Thus µ ∈ sp(p(a)) if and only ifthere is some 1 ≤ i ≤ n with λi ∈ sp(a) and we have sp(p(a)) ⊆ p(sp(a)). Now ifλ ∈ sp(a) then p(a)− p(λ) = (λ− a)b for some b ∈ A and hence is not invertible.Thus sp(p(a)) = p(sp(a)).

1.6 Exercise: Let C(z) denote the field of fractions of C[z]. Show that there isan element in C(z) which has empty spectrum.

The above shows that for a general unital algebra, it is possible for an elementto have empty spectrum. In a unital Banach algebra, however, this is not the case.This is Theorem 1.11 below. To prove it, we require a few preliminary results. Thefirst is that in a unital Banach algebra where we have a notion of convergence, wehave a theorem for a geometric series which will prove quite useful in the sequel.

1.7 Theorem: Let A be a unital Banach algebra and a ∈ A such that ‖a‖ < 1.Then 1− a is invertible and

(1− a)−1 =∞∑n=0

an.

STRUNG – INVITATION TO C∗-ALGEBRAS 6

Proof: First we note that by submultiplicativity of the norm together with theusual convergence of a geometric series, we have that∥∥∥∥∥

∞∑n=0

an

∥∥∥∥∥ ≤∞∑n=0

‖a‖n

is finite. Since A is complete, this means that∑∞

n=0 an converges to some b ∈ A.

Since

(1− a)N∑n=0

an =N∑n=0

an −N+1∑m=1

am → 1, N →∞,

we must have b = (1− a)−1, as claimed.

1.8 Lemma: Let A be a unital Banach algebra. Then Inv(A) is open in A and athe map

Inv(A)→ Inv(A) : a 7→ a−1

is differentiable.

Proof: Let a ∈ Inv(A). We will show that any b sufficiently close to a is alsoinvertible, which will show the first part of the lemma. Let b ∈ A such that‖a− b‖‖a−1‖ < 1. Then

‖ba−1 − 1‖ ≤ ‖a−1‖‖a− b‖ < 1,

so, by the previous theorem we have that ba−1 is invertible. Thus we also havethat b(a−1(ba−1)−1) = 1, so b is invertible.

To show that a 7→ a−1 is differentiable, we need to find a linear map L : A→ Asuch that, for a ∈ Inv(A),

limh→0

‖(a+ h)−1 − a−1 − L(h)‖‖h‖

= 0.

Define, for b ∈ A, L(b) = −a−1ba−1.

Let a ∈ A be invertible and let h be small enough that

‖h‖‖a−1‖ < 1/2.

Then we have that ‖a−1h‖ < 1/2 so, by Theorem 1.7 1 + a−1h is invertible and

‖(1 + a−1h)−1 − 1 + a−1h‖ = ‖∑∞

n=0(−1)n(a−1h)n − 1 + a−1h‖= ‖

∑∞n=2(−1)n(a−1h)n‖

≤∞∑n=2

‖(a−1h)‖n

≤ ‖a−1h‖2/(1− ‖a−1h‖)−1

≤ 2‖a−1h‖2.

STRUNG – INVITATION TO C∗-ALGEBRAS 7

Thus

‖(a+ h)−1 − a−1 − L(h)‖‖h‖

=‖(a+ h)−1 − a−1 + a−1ha−1)‖

‖h‖

= =‖(a(a−1a+ a−1h))−1 − (1− a−1h)a−1‖

‖h‖

≤ ‖(1 + a−1h)−1 − 1 + a−1h‖‖a−1‖‖h‖

≤ 2‖a−1‖2‖h‖2

‖h‖,

which goes to zero as h goes to zero.

In a metric space X, we will denote the open ball of radius r > 0 about a pointx ∈ X by B(x, r) and its closure by B(x, r).

1.9 Lemma: Let A be a unital Banach algebra. Then for any a ∈ A, the spectrumof a is a closed subset of B(0, ‖a‖) ⊂ C and the map

C \ sp(a)→ A : λ 7→ (a− λ)−1

is differentiable.

Proof: Once we show that sp(a) ⊂ B(0, ‖a‖), the rest follows from the previous

lemma. To prove sp(a) ⊂ B(0, ‖a‖), we need to show that if |λ| > ‖a‖ then λ− ais invertible. The details are left as an exercise.

1.10 Now we are able to prove that every element in a unital Banach algebra hasnonempty spectrum. In what follows, A∗ denotes the dual space of A, that is

A∗ = f : A→ C | f continuous and linear.A∗ can be given a topology called the weak∗-topology, which is the topology gener-ated by semi-norms of the form pa(τ) = |τ(a)| ranging over all a ∈ A. A sequence(φn)n∈N converges to φ ∈ A∗ if φn(a) → φ(a), n → ∞ for every a ∈ A (pointwiseconvergence). For further details see, for example, [1, Appendix].

1.11 Theorem: Let A be a unital Banach algbera. Then for every a ∈ A we havesp(a) 6= ∅.Proof: First of all, we may assume that a is nonzero since 0 ∈ sp(0). So let a 6= 0and assume, for contradiction, that sp(a) = ∅. We leave it as an exercise to showthat the map

C→ Inv(A) : λ 7→ (a− λ)−1

is bounded on the compact disc of radius 2‖a‖. Once this has been shown it followsthat for any φ ∈ A∗ the map

λ 7→ φ((a− λ)−1)

STRUNG – INVITATION TO C∗-ALGEBRAS 8

is also bounded. From the previous theorem, this map is also entire, which, byLiouville’s theorem, implies it must be constant. Thus φ(a−1) = φ((a − 1)−1) forevery φ ∈ A∗ leading to the contradiction that a−1 = (a− 1)−1.

The following is an immediate consequence.

1.12 Theorem: Let A be a unital Banach algebra with Inv(A) = A \ 0. ThenA = C.

1.13 The spectral radius of an element a in a unital Banach algebra A is definedto be

r(a) := supλ∈sp(a)

|λ|.

We have the following characterisation of the spectral radius which relates it tothe norm of the element a.

Theorem: For any a ∈ A we have r(a) = infn≥1 ‖an‖1/n = limn→∞ ‖an‖1/n.

Proof: Since λ ∈ sp(a) implies λn ∈ sp(an) we have |λn| ≤ ‖an‖. Thus |λ| =|λn|1/n ≤ ‖an‖1/n for every λ ∈ sp(a) and every n ≥ 1, that is,

r(a) = supλ∈sp|λ| ≤ inf

n≥1‖an‖1/n.

By definition we have that infn≥1 ‖an‖1/n ≤ lim infn→∞ ‖an‖1/n, thus we are fin-ished if we show that r(a) ≥ lim supn→∞ ‖an‖1/n.

Let D = B(0, 1/r(a)) if r(a) 6= 0 and D = C otherwise. If λ ∈ D then 1− λa isinvertible by Theorem 1.7. It follow from Lemma 1.9 that, for every φ ∈ A∗ themap

f : D → C : λ 7→ φ((1− λa)−1)

is analytic. Thus there are unique complex numbers (λn)n∈N such that

f(λ) =∞∑n=0

λnλn

whenever λ ∈ D.

But again, by applying Theorem 1.7, we have, for λ < 1/‖a‖ ≤ 1/r(a)

(1− λa)−1 =∞∑n=0

λnan.

It follows that f(λ) =∑∞

n=0 λnφ(an), so that φ(an) = λn for every n ∈ N.Thus

φ(an) → 0 as n → ∞ and therefore the sequence (φ(an))n∈N is bounded. This istrue for every φ ∈ A∗ so in fact (‖λnan‖)n∈N is also bounded by some Mλ > 0.Thus

‖an‖1/n ≤M1/nλ /|λ|,

STRUNG – INVITATION TO C∗-ALGEBRAS 9

so

lim supn→∞

‖an‖1/n ≤ 1/|λ|

for every λ ∈ D. It follows that

lim supn→∞

‖an‖1/n ≤ r(a)

as required.

1.14 Exercise: Let A be a (not necessarily unital) Banach algebra. LetA := A⊕ C as a vector space. Define a multiplication by

(a, λ) · (b, µ) = (ab+ λb+ µa, λµ),

and a norm by

‖(a, λ)‖ = ‖a‖+ |λ|Show that A is a unital Banach algebra.

Remark: When A is non unital A is called the unitisation of A. When we considerC∗-algebras in Section 3, we will have to be a little bit more careful in defining thenorm.

1.15 Exercise: Let A be a nonunital Banach algebra. The spectrum of a ∈ A isdefined to be the spectrum of a in A, that is

sp(a) = λ ∈ C | λ · 1A − a /∈ Inv(A).Give a one-line proof of Theorem 1.11 (without using Theorem 1.11!) in the casethat A is nonunital.

Exercises

1.1 Check that Examples 1.1 give Banach algebras.

1.2 Let C[z] denote the single-variable C-valued polynomials, equipped with point-wise operations and norm ‖p‖ = sup|z|=1 |p(z)|. Is this a Banach algebra?

1.3 Let A be a Banach algebra. Show that multiplication in A is continuous.

1.4 Let H be a Hilbert space with orthonormal basis (ei)i∈I . An operator T ∈B(H) is a Hilbert–Schmidt operator if

∑i∈I ‖Tei‖2 is finite. The Hilbert–Schmidt

norm ‖T‖ = (∑

i∈I ‖Tei‖2)1/2 can be defined on the set of all Hilbert–Schmidtoperators. With the usual operations for operators on a Hilbert space, are theHilbert–Schmidt operators a Banach algebra?

1.5 Let A be a unital algebra and show that the set of invertible elements Inv(A) :=a ∈ A | a is invertible in A is a group under multiplication.

1.6 What is sp(a) for a ∈ Mn and sp(f) for f ∈ C(X), where X is a compactHausdorff space?

STRUNG – INVITATION TO C∗-ALGEBRAS 10

1.7 Let H = `2(Z) = (λn)n∈Z |∑∞|n|=0 |λn|2 converges. Define the bilateral shift

operator S : `2(Z)→ `2(Z) by

S((λn)n∈Z) = (µn)n∈Z

where µn = λn−1.(a) Show that S ∈ B(H)(b) What is S∗? Is S invertible? If so, what is its inverse?(c) Show that S has no eigenvalues (i.e. for every λ ∈ C there is no ξ ∈ H suchthat Sξ = λ · ξ.) Hint: If Sξ = λ · ξ is Sξ ∈ `2?(d) Show that if |λ| = 1 then λ · 1B(H) − S is not invertible.

1.8 Let H = L2([0, 1]) = f : [0, 1] → C |∫f 2 < ∞, and consider the Banach

algebra B(H) (b) Let T ∈ B(H) be defined as

T (f)(t) =

∫ t

0

f(x)dx.

Compute the spectral radius of T . What is sp(T )?

1.9 Let X be a compact space and A a unital Banach algebra. Show that

C(X,A) := f : X → A | f continuouscan be given the structure of a Banach algebra. In the case that A = Mn we havethat C(X,A) ∼= Mn(A).

1.10 Let C(z) denote the field of fractions of C[z]. Show that there is an elementin C(z) which has empty spectrum.

1.11 Let A be a unital Banach algebra and a ∈ A. Show that sp(a) ⊂ B(0, ‖a‖)(show that if |λ| > ‖a‖ then λ− a is invertible).

1.12 Show that the map in Theorem 1.11,

C→ Inv(A) : λ 7→ (a− λ)−1

is bounded on the compact disc of radius 2‖a‖.1.13 Let A be a (not necessarily unital) Banach algebra. Let A := A ⊕ C as avector space. Define a multiplication by

(a, λ) · (b, µ) = (ab+ λb+ µa, λµ),

and a norm by

‖(a, λ)‖ = ‖a‖+ |λ|Show that A is a unital Banach algebra.

1.14 Let A be a nonunital Banach algebra. The spectrum of a ∈ A is defined tobe the spectrum of a in A, that is

sp(a) = λ ∈ C | λ · 1A − a /∈ Inv(A).

STRUNG – INVITATION TO C∗-ALGEBRAS 11

Give a one-line proof of Theorem 1.11 (without using Theorem 1.11!) in the casethat A is nonunital.

1.15 Let A be a untial Banach algebra and B ⊂ A with 1A ∈ B.

(a) Show that Inv(B) is a clopen subset of Inv(A) ∩B.

(b) Let b ∈ B. Show that spA(b) ⊂ spB(b) and ∂ spB(b) ⊂ ∂ spA(b). Showthat if C \ spA(b) has exactly one bounded component (spA(b) has no holes), thenspA(b) = spB(b).

2. The Gelfand representation

2.1 Let A be an algebra. A subset I ⊂ A is a right (left) ideal if a ∈ A and b ∈ Bthen ab ∈ B (ba ∈ B). We will call I ⊂ A an ideal if it is both a right and a leftideal. When I is an ideal, then A/I is also an algebra with the obvious definitionsfor multiplication and addition.

A/I is a unital algebra exactly when I is a modular ideal: there exists andelement u ∈ A such that a − ua ∈ I and a − au ∈ I for every a ∈ A. (What is1A/I?). Note that this implies that every ideal in a unital algebra is modular.

If A is a Banach algebra and the ideal I is norm-closed then A/I can be giventhe quotient norm

‖a+ I‖ = infb∈I‖a+ b‖, a ∈ A

making A/I into a Banach algebra.

2.2 We also have the usual notions of trivial ideals (= 0, A) and ideals generatedby a set J ⊂ A (= smallest ideal containing J). A proper ideal is one which is notequal to A (but may be zero) and a maximal ideal is a proper ideal not containedin any other proper ideal. One can use a Zorn’s Lemma argument to show thatevery proper modular ideal is contained in a maximal modular ideal. In particular,if A is unital then A every proper ideal of A is contained in a maximal ideal.

2.3 Exercise: If A is a Banach algebra then it is a proper closed maximal idealin its unitisation A (as defined in Exercise 1.14).

2.4 Proposition: Let A be a Banach algebra and I ⊂ A an ideal. If I is properand modular, than I is also proper.

Proof: Since I is modular, there is an element u ∈ A such that a− ua ∈ I anda−au ∈ I for every a ∈ A. Let b ∈ I with ‖u− b‖ < 1. Then 1−u+ b is invertibleas an element of A. Let c denote its inverse. Then

1 = c(1− u+ b) = c− cu+ cb ∈ I,

contradicting the fact that I is proper. Thus any b ∈ I must satisfy ‖u− b‖ ≥ 1.In particular, u ∈ A \ I, so I is proper.

STRUNG – INVITATION TO C∗-ALGEBRAS 12

2.5 Corollary: If I is a maximal modular ideal then it is closed.

2.6 Proposition: Let A be a unital commutative algebra and I ⊂ A a modularideal. If A is maximal then A/I is a field.

Proof: Exercise.

2.7 If A and B are Banach algebras, a map φ : A→ B is a homomorphism if it isan algebra homomorphism that is continuous with respect to the norms of A andB. If A and B are unital and φ(1A) = 1B then we call φ a unital homomorphism.The norm of a given homomorphism φ : A→ B is defined to be

‖φ‖ = sup‖φ(a)‖B | a ∈ A, ‖a‖A ≤ 1.

2.8 Exercise: Let φ : A → B be a homomorphism of Banach algebras A andB. Show that ker(φ) is a closed ideal in A.

2.9 Recall from Exercise 1.14 that A denotes the unitisation of nonunital Banachalgebra A. The map ι : A → A given by ι(a) = (a, 0) is an injective homomor-phism, so we may identify A as a subalgebra in A. We also have the canonicalprojection map π : A → C given by π((a, λ)) = λ. Its kernel is clearly A, so A isin fact an ideal in A.

2.10 Definition: Let A be a Banach algebra. A character on A is a nonzerohomomorphism τ : A→ C. Let

Ω(A) := τ : A→ C | τ a character on A

We call Ω(A) the character space of A, or based on what we’ll see below, thespectrum of A.

For commutative Banach algebras there is an important relation between charac-ters, maximal ideals, Banach algebras of the form C0(X) for some locally compactHausdorff spaces X.

2.11 Theorem: Let A be a unital commutative Banach algebra. Then

(i) τ(a) ∈ sp(a) for every τ ∈ Ω(A) and every a ∈ A,(ii) ‖τ‖ = 1, and

(iii) Ω(A) 6= ∅ and τ → ker τ is a bijection from Ω(A) to the set of maximalideals of A.

Proof: The proof is left as an exercise. For a hint, if I is a maximal ideal useTheorem 2.6 to define a homomorphism from A→ C.

Note that (ii) above says that Ω(A) is contained in the closed unit ball of thedual space A∗. Thus we may endow Ω(A) with the weak-∗ topology inherited fromA∗.

STRUNG – INVITATION TO C∗-ALGEBRAS 13

2.12 Theorem: Let A be a unital commutative Banach algebra. Then, for anya ∈ A

sp(a) = τ(a) | τ ∈ Ω(A).

Proof: Suppose that λ ∈ sp(a). The ideal generated by (a−λ) is proper since itcan’t contain 1A. It is therefore contained in some maximal ideal which is of theform ker(τ) for some τ ∈ Ω(A), in which case τ(a) = λ. Conversely, τ(τ(a)−a) = 0,so τ(a) ∈ sp(a).

The proof of this next corollary is a relatively easy exercise:

2.13 Corollary: Let A be a nonunital commutative Banach algebra. Then, forany a ∈ A,

sp(a) = τ(a) | τ ∈ Ω(A) ∪ 0.

2.14 Theorem: Let A be a commutative Banach algebra. Then Ω(A) is a locallycompact Hausdorff space. If A is unital, then Ω(A) is compact.

Proof: It follows from Theorem 2.11 that Ω(A) \ 0 is a weak-∗ subset of theclosed unit ball A∗. Thus by the Banach–Alaoglu theorem, it is compact. HenceΩ(A) is locally compact. If A is unital then one checks that in fact Ω(A) itself isclosed, hence compact.

We will now show the existence of the Gelfand representation, which says thatwe can represent any commutative Banach algebra A as functions on a locallycompact Hausdorff space which is homeomorphic to Ω(A). When we move on tothe next section, we will see that this has particularly nice consequences for C∗-algebras, in particular, it will give us a continuous functional calculus which is anindispensable tool to the theory. But first, we remain in the more general worldof Banach algebras.

2.15 Let a ∈ A and define a : Ω(A) → C by a(τ) = τ(a). Then a ∈ C0(Ω(A))(indeed, the weak-∗ topology is the coarsest topology making every a, A ∈ Acontinuous; this can be taken as its definition). The map a → a is called theGelfand transform and a is the Gelfand transform of a.

2.16 Theorem: Let A be a commutative Banach algebra with Ω(A) 6= ∅. Then

A→ C0(Ω(A)) : a 7→ a

is a norm-decreasing homomorphism and, moreover

r(a) = ‖a‖.

if A is unital and a ∈ A then sp(a) = a(Ω(A)). When A is nonunital and a ∈ A,then sp(a) = a(Ω(A)) ∪ 0.

STRUNG – INVITATION TO C∗-ALGEBRAS 14

Proof: By Theorem 2.12 and Corollary 2.13 we have r(a) = ‖a‖. Sincer(a) ≤ ‖a‖, the map is norm-decreasing. It is easy to check that it is also ahomomorphism.

2.17 Theorem: Let A be a unital Banach algebra and let a ∈ A. Let B ⊂ A bethe Banach algebra generated by a and 1A. Then B is commutative and the map

a : Ω(B)→ sp(a)

defined by

a(φ) = φ(a).

is a homeomorphism.

Proof: It is clear that B is commutative and since a is a continuous bijectionand Ω(B) is compact, it is a homeomorphism.

Exercises

2.1 By a homomorphism of Banach algebras we mean a continuous algebra ho-momorphism. Let φ : A → B be a homomorphism of Banach algebras A and B.Show that ker(φ) is a closed ideal in A.

2.2 Let I be a (not necessarily closed) ideal in a unital Banach algebra A. Showthat if I is maximal, I = I.

2.3 An ideal I in a (not necessarily unital) Banach algebra A is a modular idealif there exists and element u ∈ A such that a − ua ∈ I and a − au ∈ I for everya ∈ A.

(a) Show that every proper modular ideal is contained in a maximal modularideal.

(b) Let I be a maximal modular ideal. Show that A/I is a field.

2.4 Show that A is a maximal ideal in its unitisation A.

2.5 Prove Corollary 2.13: If A is a nonunital Banach algebra and a ∈ A then

sp(a) = τ(a) | τ ∈ Ω(A) ∪ 0.

3. C∗-algebra basics

3.1 A ∗-algebra is a C-algebra A together with an involutive (that is, an antilinearorder two isomorphism) map ∗ : A→ A. Given an element a in a ∗-algebra A, wecall a∗ the adjoint of a. An element in A is called self-adjoint if a = a∗. Anelement p ∈ A is called a projection if it is self-adjoint and p2 = p. An elementa ∈ A which commutes with its adjoint is called normal. When A is unital andu ∈ A is a normal element such that u∗u = 1 then we call u a unitary.

STRUNG – INVITATION TO C∗-ALGEBRAS 15

We will denote the set of self-adjoint elements in A by Asa and the unitaries byU(A).

3.2 Exercise: Let A be a ∗-algebra. What is sp(a∗)?

3.3 If A is a Banach algebra with involution ∗, then we call A a Banach ∗-algebraif ‖a∗‖ = ‖a‖ for every a ∈ A.

3.4 An abstract C∗-algebra is Banach ∗-algebra whose norm satisfies the followingequality, called the C∗-equality

‖a∗a‖ = ‖a‖2 for every a ∈ A.

We often call a norm which satisfies the C∗-equality a C∗-norm.

What may appear to be only a minor requirement for the norm in fact givesa C∗-algebra many nice structural properties that we don’t see in an arbitraryBanach algebra. First, let’s look at a few examples.

3.5 Example: Let n ∈ N and Mn := Mn(C) be the set of n × n matri-ces with complex entries. Equip Mn with the operator norm, that is, ‖A‖ =supx∈Cn,‖x‖=1 |A(x)|. Then Mn is a C∗-algebra under the usual matrix multiplica-tion and addition and with involution given by taking adjoints.

More generally, if H is a Hilbert space then B(H) is a C∗-algebra.

3.6 Example: Let H be a Hilbert space and let K(H) denote the subalgebra ofcompact operators. Then K(H) is also a C∗-algebra with the inherited structurefrom B(H).

More generally, if A is any closed self-adjoint subalgebra of B(H), then A is alsoa C∗-algebra with the inherited structure. Such a C∗-algebra is called a concreteC∗-algebra.

3.7 Example: Let X be a locally compact Hausdorff space. We say that afunction f : X → C vanishes at infinity if, for every ε > 0 there is a compact setK ⊂ X such that |f(x)| < ε for every x ∈ X \ K. Let C0(X) = f : X → C |f vanishes at infinity, and equip C0(X) with pointwise operations, sup norm and

for f ∈ C0(X), define f ∗(x) := f(x). Then C0(X) is a C∗-algebra which is unitalif and only if X is compact. In this case we denote C0(X) by C(X).

3.8 Example: By an ideal I in a C∗-algebra A, we will mean a self-adjoint, closedtwo-sided ideal. Given such an ideal, we may define the quotient C∗-algebra A/Ias we did for a Banach algebra together with the obvious involution induced fromA.

A few nice properties that we obtain from having a C∗-norm can be noticedstraight away. For example, it is automatic in a unital C∗-algebra we have ‖1‖ = 1.More generally, if u is a unitary, ‖u‖ = 1 and also if p is a projection then ‖p‖ = 1.

STRUNG – INVITATION TO C∗-ALGEBRAS 16

3.9 This gives us information about the spectrum of a unitary u: Using Lemma 1.9,if λ ∈ sp(u) then |λ| ≤ ‖u‖ = 1. Since u is invertible, we must also have λ−1 ∈sp(u−1) = sp(u∗) ≤ ‖u∗‖ = 1. Thus |λ| = 1 so spu is a closed subset of T.

3.10 With a bit more, work, we can also show that for any a ∈ Asa we have sp(a) ⊂R. First, note that for any element a in a unital Banach algebra

∑∞n=0 ‖an/n!‖ ≤∑∞

n=0 ‖a‖n/n! and so converges. We set ea :=∑∞

n=0 an/n!. For any a ∈ A, one

can check that the map

φ : R→ A : t→ eta

is differentiable at every t ∈ R with derivative aφ(t) and it satisfies φ(0) = e0 = 1.Moreover, these completely characterise t → eta in the sense that if ψ is anotherfunction with these properties then necessarily φ = ψ. (Details: exercise; recallhow this is done in the case that A = R).

Using this characterisation, it follows that ea+b = eaeb for any a, b ∈ A withab = ba. In particular, ea is always invertible.

3.11 Now let A be a unital C∗-algebra and If a is self-adjoint, then eia is invertibleof norm 1 hence it is a unitary. As we saw in 3.9, we must then have sp(eia) ⊂ T.Let λ ∈ sp(a). Let b =

∑∞n=2 i

n(a− λ)n−1/n!. Note that b commutes with a. Wehave

eia − eiλ = (ei(a−λ) − 1)eiλ = (a− λ)beiλ.

Since b commutes with a and hence (a − λ) and (a − λ) is not invertible, we seethat eia − eiλ is not invertible. Thus eiλ ∈ sp(eia) ⊂ T so we must have λ ∈ R.

3.12 Theorem: Let A be a C∗-algebra and let a ∈ Asa. Then r(a) = ‖a‖.Proof: Exercise.

Note that this means that the norm of an element in a in a C∗-algebra A dependsonly on spectral data since: if a ∈ A then a∗a is self-adjoint and

‖a‖ = ‖a∗a‖1/2 = (r(a))1/ =

(sup

λ∈sp(a)

|λ|

)1/2

.

This gives us the next theorem.

3.13 Theorem: If the ∗-algebra A admits a norm which makes A into a C∗-algebra, then it is the unique C-norm on A.

Minimal unitisation. We saw in Exercise 1.14 that a nonunital Banach algebraA can be embedded into a unital Banach algebra A. There is also a way of defininga unitisation (in fact, more than one) of a nonunital C∗-algebra. Unfortunately,we can’t simply take A⊕C with multiplication and norm as given in Exercise 1.14.The reason is that the norm there is not a C∗-norm (check!). So we have to be abit more careful in how we adjoin a unit to a nonunital C∗-algebra.

STRUNG – INVITATION TO C∗-ALGEBRAS 17

3.14 If T : A→ B is a linear operator, we equip it with the operator norm:

‖T‖ := supa∈A,‖a‖≤1

‖T (a)‖B,

which is just the usual norm for a linear operator if we consider A and B as Banachspaces.

3.15 A left multiplier L of A is a bounded linear operator L : A→ A which satisfiesL(ab) = L(a)b for every a, b ∈ A. If a ∈ A, then a can act on A as a left multiplierb 7→ ab. Similarly, it is easy to see that any λ ∈ C induces a left multiplier on Aby b 7→ ab.

3.16 Let A = A⊕C as a vector space. We endow A with the same multiplicationas for the Banach algebra unitisation of A. Define an involution ∗ : A → A by(a, λ)∗ = (a∗, λ)). Now, to make A into a C∗-algebra , we view the elements of Aas left operators on the A.

‖(a, λ)‖A := supb∈A,‖b‖≤1

‖ab+ λb‖A.

One then checks that this makes A into a unital C∗-algebra. Moreover, themap a 7→ (a, 0) identifies as an ideal in A. We call A the minimal unitisation ofA. What might be thought of as the “maximal” unitisation of A, the multiplieralgebra, is discussed in Section 8. Note that this norm is the unique norm makingA into a C∗-algebra and unless otherwise specified, this is the norm we use (ratherthan the one defined in Exercise 1.14) for A.

3.17 Example: Let A = C0(X) where X is a locally compact Hausdorff space.Then A = C(X ∪ ∞) where X ∪ ∞ is just the one point compactification ofX []. In this way, we think of adjoining a unit as the noncommutative version ofcompactification.

The nice thing about the unitisation is that we will now be able to prove manytheorems in the (usually easier) unital setting without any loss of generality.

3.18 Let A and B be C∗-algebras. A ∗-homomorphism φ : A → B is an algebrahomomorphism which is involution-preserving, φ(a∗) = φ(a)∗ for every a ∈ A. IfA and B are unital, then, as before, we say that φ is a unital ∗-homomorphism ifφ(1A) = 1B.

If B is unital and φ : A → B is a ∗-homomorphism then there exists a uniqueextension φ : A→ B such that φ is unital.

3.19 The following proposition is another nice implication of the C∗-equality.

Proposition: A ∗-homomorphism between C∗-algebras is always norm-decreasing. In particular, it is always continuous.

STRUNG – INVITATION TO C∗-ALGEBRAS 18

Proof: Let φ : A→ B be a ∗-homomorphism, and let a ∈ A. As we saw in 3.18,by replacing A and B by their unitisations if necessary, we can assume that φ, Aand B are all unital. Since φ(1A) = 1B, it is easy to check that if a is invertible sois φ(a) is invertible in B. It follows that sp(φ(a)) ⊂ sp(a). The result now followsfrom Theorem 3.12.

3.20 Corollary: Any ∗-isomorphism of C∗-algebras is automatically isometric.

3.21 Recall that Ω(A) is the spectrum, or character space, of A (see 2.10). If X islocally compact and Hausdorff and x ∈ X then the evx(f) = f(x) is a characterof C0(X). In fact, this is all of them.

Theorem: Let A = C(X) for some compact Hausdorff space X. Then

Ω : X → Ω(A) : x→ evx

is a homeomorphism.

Proof: Let (xλ)Λ is a net in X with limλ xλ → x ∈ X. Then evxλ(f) = f(xλ)→f(x) − evx for every f ∈ C(X), so (evλ)Λ is weak-∗ convergent to evx. Thus themap is continuous.

Suppose now that x 6= y ∈ X. Then by Urysohn’s lemma, there is f ∈ C(X)such that f(x) = 1 and f(y) = 0. Thus evx 6= evy and so we see that the map isinjective.

Now let us prove surjective:. Let τ be a character on C(X). Let M := ker(τ);this is a maximal, hence proper, ideal in C(X). We show that M separates points.If x 6= y the there is f ∈ C(X) with f(x) = 1 and f(y) = 0. Now f − τ(f) ∈ Msatisfies f(x)− τ(f) 6= f(y)− τ(f) so by the Stone–Weierstrass theorem, there isx ∈ X such that f(x) = 0 for every f ∈M .

Thus (f − τ(f))(x) = 0 and so f(x) = τ(f) for every f ∈ C(X). It follows thatevx = τ and so Ω is surjective.

Since any continuous bijective map from a compact space is a homeomorphism,the result follows.

3.22 Now we come to one of the most important results in C∗-algebra theory, thatfor a commutative C∗-algebra , the Gelfand Transform of 2.15 is an isometric ∗-isomorphism. This gives a complete characterisation of commutative C∗-algebras :they are always, up to ∗-isomorphism, of the form C0(X)) for some locally compactHausdorff space X.

Theorem: [Gelfand–Naimark] Let A be a commutative C∗-algebra . Then theGelfand transform

Γ : A→ C0(Ω(A)) : a→ a

is an isometric ∗-isomorphism.

STRUNG – INVITATION TO C∗-ALGEBRAS 19

Proof: If φ ∈ Ω(A) then φ(a) ∈ R whenever a ∈ Asa. Thus for any A 3 c = a+ib

with a, b ∈ Asa we have φ((a + ib)∗) = φ(a − ib) = φ(a) − iφ(b) = (φ(a) + iφ(b)),

that is, φ is a ∗-homomorphism. It follows that a∗(φ) = (a)(φ) for any a ∈ A andany φ ∈ C0(Ω(A)) meaning Γ is a ∗-homomorphism. This moreover implies that

‖a‖2 = ‖a∗a‖ = ‖a∗a‖ = r(a∗a) = ‖a∗a‖ = ‖a‖2, so the map is isometric. Finally,to see that the map is also surjective, we appeal to the Stone–Weierstrass theorem:the image of Γ contains functions which don’t simultaneously vanish any where onΩ(A) and also separates points, thus is exactly C0(Ω(A)).

At first glance, Theorem 3.22 applies to the relatively small class of commutativeC∗-algebras. While it is true that in greater generality we don’t have such anexplicit characterisation for a class of C∗-algebras, what we do get is an extremelyuseful tool: the continuous functional calculus for normal elements.

3.23 Theorem: Let A be a unital C∗-algebra and let a ∈ A be a normal element.Then there is a map

γ : C(sp(a))→ A : (z 7→ z)→ a

is an isometric ∗-homomorphism and γ(C(sp(a))) = C∗(a, 1)

Proof: Since a is a normal element, C∗(a, 1) is abelian. Thus by Theorem 3.22we have a ∗-isomorphism

Γ : C∗(a, 1)→ C(Ω(C∗(a, 1))) : a 7→ a.

By Theorem 2.17, h : Ω(C∗(a, 1)) → sp(a) is a homeomorphism so we also havean isomorphism ψ : C(sp(a)) → C(Ω(C∗(a, 1))). Let f(z) = z for z ∈ sp(a).Let γ = Γ−1 ψ Since C(sp(a)) is generated by 1 and f , γ is the unique unital∗-homorphism with γ(f) = a. Clearly γ is isometric and its image is C∗(a.1).

3.24 Exercise: Write down and prove the nonunital version of Theorem 3.23.

The reason for highlighting the unital version is the following categorical theorem:

3.25 Theorem: The correspondence between X and C(X) is a categorical equiv-alence between the category of compact Hausdorff spaces and continuous maps tothe category of unital C∗-algebras and unital ∗-homomorphisms.

3.26 This is where we get the nomer “noncommutative topology” for the study ofC∗-algebras. In general it useful to think of the C∗-landscape as having two coastsat opposite ends, one of which consists of bounded operators on Hilbert spaces andmatrix algebras, the other consisting of commutative C∗-algebras . The interestingpart of the theorem comes as we move inland, as most C∗-algebras , of course, liesomewhere in between. Some of our best tools are brought in via either coast andit’s often useful to keep these two examples in mind.

With this result in hand, we will be able to do a lot more in our C∗-algebras. If p ∈ C[z1, z2] is a polynomial, then since A is an algebra, it is clear that

STRUNG – INVITATION TO C∗-ALGEBRAS 20

p(a, a∗) ∈ C∗(a, 1). (In the nonunital case we would require that p have no constantterms; then p(a, a∗) ∈ C∗(a, 1).) Since p of this form are dense in C(sp(a)), we canthen define Using Theorem 3.23 we define

f(a) := γ(f) ∈ C∗(a, 1),

or when A is nonunital, for f ∈ C0(sp(a)),

f(a) := γ(f) ∈ C∗(a).

The following is sometimes called the spectral mapping theorem.

3.27 Theorem: Let A be a C∗-algebra and let a ∈ A be a normal element. Thenfor any f ∈ C0(sp(a)), the element f(a) is normal and we have

sp(f(a)) = f(sp(a)).

Furthermore, if g ∈ C0(sp(f(a)) then g f(a) = g(f(a)).

Proof: Exercise.

3.28 Exercise: Let A be a ∗-algebra. Two elements a, b ∈ A are said to beunitarily equivalent if there exists a unitary u ∈ A such that uau∗ = b. Show thatif a and b are unitarily equivalent then sp(a) = sp(b).

Exercises

3.1 Let A be a ∗-algebra and let a ∈ A. Describe sp(a∗).

3.2 Let A be a C∗-algebra and p ∈ A a nonzero projection. What is sp(p)?

3.3 Let ψ : A → R be a differentiable map with derivative aψ(t) and ψ(0) = 1.Show that ψ(t) = eta.

3.4 A C∗-algebra is simple if it has no nontrivial (closed, two-sided) ideals. Givean example of a finite-dimensional simple C∗-algebra. Show that the compactoperators K are a simple C∗-algebra. (Here K denotes compact operators on aseparable Hilbert space. Since there is only one separable Hilbert space up toisomorphism we often don’t reference the underlying Hilbert space).

3.5 Let A = C(X) where X is a compact Hausdorff space. Show that if F ⊂ X isa closed subset then

f ∈ A | f |F = 0is an ideal. Show that every ideal in A has this form.

3.6 Describe all simple commutative C∗-algebras.

3.7 Let A ⊂ B(H) be a concrete C∗-algebra. Show that Mn(A) admits a C∗-normand is a C∗-algebra (ie. complete with respect to this norm).

3.8 Let A be a C∗-algebra and let a ∈ Asa. Then r(a) = ‖a‖. Show that if A is a∗-algebra admitting a C∗-norm, then the norm is the unique C∗-norm on A.

STRUNG – INVITATION TO C∗-ALGEBRAS 21

3.9 Let A be a unital C∗-algebra. Suppose a, b ∈ A are normal elements that areunitarily equivalent (that is, there exists a unitary u ∈ U(A) such that u∗au = b).Show that the C∗-subalgebras C∗(a, 1) and C∗(b, 1) are isomorphic.

3.10 Let A be a unital algebra. Let a ∈ Asa and 0 < ε < 1/4. Suppose sp(a) ⊂[0, ε] ∪ [1− ε, 1]. Show that there is a projection p ∈ A with ‖p− a‖ ≤ ε.

3.11 Write down and prove the nonunital version of Theorem 3.23.

3.12 Let ι : (0, 1) →R be the inclusion map. Then ι is continuous. Show thatι∗ : C0(R) → C0((0, 1) defined by ι∗(f) = f ι is not a ∗-homomorphism. Whatgoes wrong?

3.13 Prove the spectral mapping theorem: Let A be a C∗-algebra and let a ∈ Abe a normal element. Then for any f ∈ C0(sp(a)), the element f(a) is normal andwe have

sp(f(a)) = f(sp(a)).

Furthermore, if g ∈ C0(sp(f(a)) then g f(a) = g(f(a)).

4. Positive elements

We’ll give the functional calculus a workout in this section. The first thing is todefine positive elements in a C∗-algebra as well as a partial order on its self-adjointelements.

4.1 Exercise: Let A be a ∗-algebra. Show that any element a ∈ A can bewritten as a = a1 + ia2 where a1, a2 are self-adjoint. In this sense, the self-adjointelements play the role of “real” elements in A, in analogy to real numbers in C.

Since we’re talking about self-adjoint elements, it might be useful to considerwhat is going on in the case of functions defined on an interval in R.

4.2 Let A be a C∗-algebra. An element a ∈ A is said to be positive if a ∈ Asa andsp(a) ⊂ [0,∞). The set of positive elements is denoted A+.

If a, b ∈ Asa then we write a ≤ b if b− a is positive. This defines a partial orderon Asa.

4.3 Lemma: Let X be a locally compact Hausdorff space and let f ∈ C0(X) satisfyf(x) ≥ 0 for all x ∈ X. Then f has a unique positive square root.

4.4 The following fact often proves handy and makes use of the functional calculus.

Proposition: Every positive element has a unique positive square root.

Proof: Let a ∈ A+. It is clear this holds if a = 0, so assume otherwise. Sincesp(a) ⊂ [0,∞) there is a function f ∈ C(sp(a)), f ≥ 0 which satisfies (f(t))2 = tfor every t ∈ sp(a). Since a is normal, we can set a1/2 := f(a), which is a positivesquare root for a ∈ A. Suppose there is another b ∈ A+ with b2 = a. Since b 6= 0

STRUNG – INVITATION TO C∗-ALGEBRAS 22

and b2 commutes with a, so does b, and by approximating by polynomials, b alsocommutes with a1/2. Thus the C∗-algebra B generated by a and b is commutative,and moreover sp(a) ⊂ Ω(B). Thus by uniqueness of the square-root in C(Ω),b = a1/2.

4.5 Proposition: A unital C∗-algebra is linearly spanned by its unitaries.

Proof: We saw that every element can be written as a linear combination ofself-adjoint elements, so we need only show that the self-adjoints are spannedby unitaries. Let a ∈ Asa. By scaling if necessary, we may also assume that‖a‖ ≤ 1. ln this case, 1 − a2 ≥ 0 and so has a positive square root,

√1− a2.

Let u1 = a − i√

1− a2 and u2 = a + i√

1− a2; it is easily checked that these areunitaires and that

a = u1/2 + u2/2.

4.6 Recall that a convex cone is a subset of a vector space which is closed underlinear combinations with positive coefficients.

Theorem: The A+ is closed convex cone.

Proof: Without loss of generality, we may assume that A is unital. Let a ∈ A+

and λ ≥ 0. Then clearly λa ≥ 0 so we only need to check that the sum of twopositive elements is again positive. Let a, b ∈ A+. We claim that if f ∈ C(X)where X is a compact subset of R that f is positive if there is r ∈ R≥0 such that‖f − r‖ ≤ r. If r = 0 we must have f = 0, which is positive. Suppose thatr > 0 but f is not positive. Then there is some t ∈ X such that f(t) < 0. Then|f(t)− r| > r so ‖f − r‖ = supt∈X |f(t)− r| > r, a contradiction. This proves theclaim.

Note that furthermore, if f is positive then ‖f − r‖ < r if r ≤ ‖f‖Since a + b is self-adjoint, so we may identify C∗(a + b, 1) ∼= C(sp(a + b)) and

a+ b with f(t) = t. Thus we need only find such an r satisfying ‖a+ b− r‖ ≤ r.Let r = ‖a‖ + ‖b‖. We have that ‖a − ‖a‖‖ ≤ ‖a‖ and ‖b − ‖b‖‖ ≤ ‖b‖ by theabove. Thus

‖a+ b− ‖a‖ − ‖b‖‖ = ‖a− ‖a‖+ b− ‖b‖‖ ≤ ‖a− ‖a‖‖+ ‖b− ‖b‖ ≤ ‖a‖+ ‖b‖.Hence a+ b is positive.

To show that it is closed, first note that by the above

B := a ∈ A+ | ‖a‖ ≤ 1 = a | ‖a− 1‖ ≤ 1 ∩ Asa.Both a | ‖a− 1‖ ≤ 1 and Asa are closed, thus RB = RA+ = A+ is closed.

4.7 Let a ∈ Asa. Let a− ∈ C(sp(a)) be the function a(t) = −t when t ≤ 0 and 0otherwise. Denote by a+ ∈ C(sp(a)) the function a+(t) = t for every t ≥ 0 and 0otherwise. Then a+, a− ∈ A+ and a = a+ − a−. Note also that a+a− = 0.

STRUNG – INVITATION TO C∗-ALGEBRAS 23

The above observation also means that a+ − a = a− ∈ A+ and a ≤ a+. This,together with the previous theorem, gives us the next corollary.

Corollary: (Asa,≤) is upwards directed.

4.8 Theorem: For every a ∈ A, the element a∗a is positive.

Proof: Clearly a∗a ∈ Asa. Suppose that −a∗a ∈ A+. Then by 1.4, so is −aa∗.We have a = b + ic for some b, c ∈ Asa, so a∗a = b2 + ibc − icb + c2 and aa∗ =b2 − ibc + icb + c2. Hence a∗a = 2b2 + 2c2 − aa∗ is the sum positive elements andmust also be positive. Thus ‖a∗a‖ = 0, hence ‖a‖2 = 0 and therefore a = 0.

Suppose that a 6= 0. Then by the above, −a∗a is not positive. As we saw in 4.7,we can write a∗a = b−c where b, c ∈ A+ and bc = 0. We will use the above to showthat c = 0. Consider ac. We have −(ac)∗(ac) = −ca∗ac = −c(b − c)c = c3 ∈ A+,hence ac = 0 and finally c = 0.

The next theorem gives some more useful facts about the self-adjoint and thepositive elements.

4.9 Theorem: Let A be a C∗-algebra . Let a, b ∈ Asa with a ≤ b. Then

(i) for any c ∈ A, we have c∗ac ≤ c∗bc;(ii) ‖a‖ ≤ ‖b‖;(iii) if A is unital and a, b ∈ Inv(A) then b−1 ≤ a−1.

Proof:

(i) Since a ≤ b, the element b − a ∈ A+ and therefore has a positive squarerooot. Then c∗bc− c∗ac = c∗(b− a)c = ((a− b)1/2c)∗((a− b)1/2c), which ispositive by Proposition 4.8.

(ii) Without loss of generality, we may assume that A is unital. Then, usingthe functional calculus, we have b ≤ ‖b‖, so also a ≤ ‖b‖. Then, sinceC∗(a, 1)→ C(sp(a)) is isometric, we have that ‖a‖ ≤ ‖b‖.

(iii) For any a ∈ A, if c ∈ Inv(A) then λ ∈ sp(c) if and only if λ−1 ∈ sp(c)so a−1, b−1 ∈ Asa. By (i), b−1/2ab−1/2 ≤ b−1/2bb−1/2 = 1. So, using (ii),‖b−1/2a1/2‖2 < 1. Thus ‖a1/2b−1a1/2‖ ≤ 1 and a1/2b−1a1/2 ≤ 1. By (i)again, mulitiplying a−1/2 on either side gives b−1 < a−1.

We also have the next theorem, which we include separately because the proofis a little trickier.

4.10 Theorem: Let A be a C∗-algebra and let a, b ∈ A+ with a ≤ b. Thenan ≤ bn for all 0 < n ≤ 1.

STRUNG – INVITATION TO C∗-ALGEBRAS 24

Proof: Let a, b ∈ A+ with a ≤ b. Let ε > 0 and put c = b + 1Aε. Then a ≤ c

and c is invertible in A. Set

S = n ∈ (0,∞) | an ≤ cn.

Note that 1 ∈ S. Suppose that (tn)n∈N is a sequence in S converging to t ∈ (0,∞).Then limn→∞ a

tn − ctn = at − ct ≥ 0 since A+ is closed.

Since c ≥ 0, so is c−1 and thus has a positive square root. By Theorem 4.9 (i),this implies that c−1/2ac−1/2 ≤ 1A, and, using the C∗-equality that ‖a1/2c−1/2‖ ≤ 1.We saw earlier that sp(xy) \ 0 = sp(yx) \ 0 for any x, y ∈ A. Thus

‖c−1/4a1/2c−1/4‖ = r(c−1/4a1/2c−1/4) = r(a1/2c−1/2) ≤ 1,

so c−1/4a1/2c−1/4 ≤ 1 and a1/2 ≤ c1/2. Thus 1/2 ∈ S.

One shows similarly that if s, t ∈ S then (s+ t)/2 = z ∈ S. Indeed,

‖c−z/2azc−z/2‖ = r(c−z/2azcz/2)

= r(c−s/2as/2at/2c−t/2)

= ‖(c−s/2as/2)(at/2c−t/2)‖≤ ‖c−s/2as/2‖‖at/2c−t/2‖≤ 1.

Thus (0, 1] ⊂ S. Since ε was arbitrary, we have that an ≤ bn for every n ∈ (0, 1].

4.11 What happens for n ≥ 1? If A = C(X) is commutative and f, g ≥ 0 withf ≤ g, then certainly f 2 ≤ g2. In this case, however, an attempt to look to thecommutative algebras leads us astray. Let us consider things from the other side,then, in spirit of 3.26. In fact, we need only consider M2 to see that things can gowrong. The conclusion of the last theorem no longer necessarily hold. Let

a =

(1 00 0

)and b =

(1/2 1/21/2 1/2

)Then a ≤ a+ b but

(a+ b)2 − a2 = ab+ ba+ b2 = a+ 2ab+ b =

(3/2 11 1/2

)is not positive.

By the way, this above example is also a good illustration of the following: It isoften very useful to look to the 2×2 matrices as a test case when deciding whetheror not something holds in a C∗-algebra.

STRUNG – INVITATION TO C∗-ALGEBRAS 25

Approximate units and hereditary C∗-subalgebras. We have already seenthat it is always possible to adjoin a unit to a nonunital C∗-algebra. In many cases,however, this is not necessarily useful. For example, if one is investigating simpleC∗-algebras (no nontrivial ideas), then attaching a unit destroys simplicity. Wecan, however, often find an approximate unit. Let’s consider continuous functionson some open interval I in R, for example. Then the C∗-algebra C0(I) is of coursenonunital. However, for any finite subset of functions F ⊂ C0(I) nad any ε > 0,we can always find another function, g such that ‖gf − f‖ < for every f ∈ F .Since C0(I) is separable, one can find a nested increasing sequence of these finitesubsets F0 ⊂ F1 ⊂ · · · that exhausts the algebra. Then, for every n ∈ N find gnsuch that fgn = f for every f ∈ Fn. This gives us a sequence (gn)N∈N ⊂ C0(I)which itself will not converge, but which still satisfies

limn→∞

gnf = f for every f ∈ C0(I).

In the completely noncommutative case, consider the compact operators K(H)on a separable Hilbert space H with orthonormal (ei)i∈N. They are of coursenonunital but, thinking of K as infinite matrices, then if pn is the projection∑n

i=1 ei we also get a sequence that won’t converge, but, for any a ∈ K we have

limn→∞

pna = limn→∞

apn = a.

4.12 Definition: Let A be a C∗-algebra . An approximate unit for A is anincreasing net (uλ)λ∈Λ of positive elements such that

limλuλa = lim

λauλ = a,

for every a ∈ A.

4.13 Lemma: Let A be a C∗-algebra and let A1+ := a ∈ A | a ∈ A+, ‖a‖ < 1.

Then the set of (A1+,≤) is upwards-directed.

Proof: First, let a, b ∈ A+. Then (1 + a), (1 + b) ∈ Inv(A). Suppose that a ≤ b.Then 1 + a ≤ 1 + b so (1 + b)−1 ≤ (1 + a)−1 and a(1 + a)−1 = 1 − (1 + a)−1 ≤1− (1 + b)−1 = b(1 + b)−1. Note that a(1 + a)−1 and b(1 + b)−1 are both in A1

+.

Now suppose that a, b ∈ A1+. Let x = a(1 − a)−1 and y = b(1 − b)−1 and set

c = (x + y)(1 + x + y)−1. Since x ≤ x + y we have that a = x(x + 1)−1 ≤ c.Similarly b ≤ c.

4.14 Theorem: Every C∗-algebra A has an approximate unit. If A is separable,then A has a countable approximate unit.

Proof: Let Λ be the upwards directed set (A1+,≤) and put uλ = λ for each λ ∈ Λ.

Then (uλ)Λ is an increasing net of positive elements with norm less than 1. Wemust show that limλ uλa = limλ auλ = a for every a ∈ A. It is enough to show

STRUNG – INVITATION TO C∗-ALGEBRAS 26

that this holds when a ∈ A+. Let ε > 0. Let Γ : C∗(a)→ C0(X)) be the Gelfandtransform. Let f = Γ(a) and let K = x ∈ X|f(x)| ≥ ε.

Let δ > 0 such that δ < 1 and 1 − δ < ε. Let gδ(x) = δ if x ∈ K andvanishing outside K (this is possible since K is compact). Then gδ ∈ C0(X)1

+

and ‖gδf − f‖ < ε. Since Γ is isometric, Γ−1(gδ) = µ for some µ ∈ Λ and‖uµa− a‖ = ‖auµ − a‖ < ε.

Suppose that λ ∈ Λ satisfies µ ≤ λ. Then 1− uλ ≥ 1− uµ so a1/2(1− uλ)a1/2 ≤a1/2(1 − uµ)a1/2 and hence ‖a1/2(1 − uλ)a1/2‖ ≤ ‖a1/2(1 − uµ)a1/2‖ and applyingthe C∗-equality, ‖a− auλ‖ < ‖a− auµ‖ < ε. Hence limλ auλ = limλ uλ = a.

The proof that if A is separable then A admits a countable approximate unit isan exercise.

The following is a very useful theorem. We omit the proof for now because ituses some techniques that have not yet been described, but will return to it in thenext section.

4.15 Theorem: Let A be a C∗-algebra and I a closed ideal in A. Then I hasa quasicentral approximate unit for A, that is, I has an approximate unit (uλ)λ∈Λ

satisfying

‖uλa− auλ‖ → 0 as λ→∞

for every a ∈ A.

4.16 We have already defined C∗-subalgebras as well as ideals. There is anotherimportant substructure in C∗-algebras coming from the order structure.

Definition: A C∗-subalgebra B ⊂ A is called hereditary if, whenever b ∈ B anda ≤ b, then a ∈ B.

4.17 This next example is called a corner of a C∗-algebra .

Theorem: Let p ∈ A be a projection. Then pAp = pap | a ∈ A is a hereditarysubalgebra of A.

Proof: Since p is a projection, it is easy to check that pAp is C∗-subalgebra. Wemay assume that a ∈ pAp is positive since if b ≤ a then b ≤ a+. Suppose first thatb ≤ a with b also positive or b = −c where c is positive. Then (1 − p)b(1 − p) ≤(1 − p)a(1 − p) = 0. So ‖(1 − p)b(1 − p)‖ = ‖(1 − p)b1/2‖2 and we see thatpb = pb1/2b1/2 = b1/2b1/2 = b. Similarly, bp = b, hence b ∈ pAp. If b = b+− b− with

b+, b− positive and b+b− = b−b+ = 0. Since b ≤ a we have that b2+ ≤ b

1/2+ ab

1/2+ and

b2+/‖b‖ ≤ (b

1/2+ /‖|b1/2‖)a(b

1/2+ /‖b1/2‖) ≤ a. Since b2

+/‖b‖ is positive, b2+/‖b‖ ∈ pAp.

Hence b+ ∈ pAp. Then since −b− ≤ b ≤ a ∈ pAp, we also have b− ∈ pAp and sob ∈ pAp, as required.

For a positive element, we have the following generalisation:

STRUNG – INVITATION TO C∗-ALGEBRAS 27

4.18 Theorem: Let a ∈ A+. Then a ∈ aAa and aAa is the hereditary C-subalgebra generated by a. If B is a separable hereditary C∗-algebra, then B = aAafor some a ∈ A+.

Proof: It is clear that aAa is a C∗-subalgebra. Let (uλ)Λ be an approximate unitfor A. Then limλ auλa = a2 so a2 ∈ aAa. Thus C∗(a2) ⊂ aAa and by uniquenessof the positive square root, also a ∈ aAa.

The proof that aAa is hereditary is similar to the case for a corner, so is left asan exercise.

Now suppose that B ⊂ A is hereditary and separable. Since B is separable, itcontains a countable approximate unit, say (un)n∈N. Let a =

∑∞n=1 2−nun. Then

a ∈ B and a ≥ 0. Thus aAa ⊂ B. For each N \ 0 we have that 2−nun ≤ a soun ∈ aAa. Thus, if b ∈ B we have b = limn→∞ unbun where each unbun ∈ aAahence b ∈ aAa and so we have shown that B = aAa.

4.19 Lemma: Let L be a closed left ideal of A. Then L has a left approximateunit, that is, (uλ)λ∈Λ ⊂ I with

limλauλ = a

for every a ∈ A.

Proof: Observe that L ∩ L∗ is a C∗-subalgebra of A and thus by Theorem 4.14has an approximate unit (uλ)λ∈Λ. Let a ∈ L. Then a∗a ∈ L ∩ L∗. Since

‖a− auλ‖2 = ‖a∗a− a∗auλ − uλa∗a+ uλa∗auλ‖,

we have

‖a− limλauλ‖2 = 0.

Thus a = limλ auλ.

4.20 Theorem: Let A be a C∗-algebra . There is a one-to-one correspondencebetween closed left ideals of A and hereditary subalgebras of A given by

I 7→ I∗ ∩ I; B 7→ a ∈ A | a∗a ∈ B.

Proof: Let I be a closed left ideal in A and suppose b ∈ (I ∩ I∗)+. Since I isa closed left ideal, it contains a left approximate unit (uλ). If a ∈ A+ satisifiesa ≤ b, then (1− uλ)a(1− uλ) ≤ (1− uλ)b(1− uλ). Then

‖a1/2(1− uλ)‖2 = ‖(1− uλ)a(1− uλ)‖≤ ‖(1− uλ)b(1− uλ)‖= ‖b1/2 − b1/2uλ‖2

→ 0.

STRUNG – INVITATION TO C∗-ALGEBRAS 28

Thus ‖a1/2(1 − uλ)‖2 → 0 and so a1/2 is a limit of elements in I. It follows thata1/2, and hence a ∈ I. Thus I∗ ∩ I is a hereditary C∗-subalgebra of A.

Now suppose that B is a hereditary C∗-subalgebra. Let I = a ∈ A | a∗a ∈ B.Let a ∈ I, b ∈ A and without loss of generality, assume ‖b‖ ≤ 1. Then (ab)∗(ab) =b∗a∗ab ≤ a∗a, so ab ∈ I since B is hereditary. It is clear that I is closed since Bis; thus I is a closed left ideal.

Finally, it is easy to check that the maps are mutual inverses.

4.21 Corollary: Every closed ideal is hereditary.

Proof: Let I ⊂ A be a closed ideal. Then in particular I is a closed left idealand we have I = I ∩ I∗ since if a ∈ I and (uλ)Λ is a left approximate unit for Ithen a∗ = limλ(auλ)

∗ = limλ uλa∗ ∈ I.

A hereditary C∗-subalgebra often inherits properties of the C∗-algebra itself. Weshall see more examples later, but for now we can show the following

4.22 Theorem: If A is a simple C∗-algebra, then so is every hereditary subalgebraB ⊂ A.

Proof: We claim that every ideal J ⊂ B is of the form I ∩ B for some idealI ⊂ A. Suppose that I is an ideal in A and let b ∈ I ∩ B. Clearly I ∩ B is closedunder addition and scalar multiplication and is norm-closed. Let c ∈ B. Withoutloss of generality assume that ‖c‖ ≤ 1 and that c ≥ 0. In that case cb ≤ b andsimilarly bc ≤ b. Thus cb, bc ∈ I ∩B.

Now suppose that J ⊂ B is an ideal. Let (uλ)Λ be a quasicentral approximateunit for J . Set I = auλ | a ∈ A, λ ∈ Λ. Then J = I ∩B; the details are easy tocheck.

Exercises

Let A be a C∗-algebra

4.1 A partial isometry in a C∗-algebra is an element satisfying v = vv∗v andv∗ = v∗vv∗.

(a) Show that v∗v and vv∗ are projections.

(b) Let A = Mn and let p, q be projections. If tr(p) ≤ tr(q) show that there isa partial isometry v ∈ Mn such that v∗v = p and vv∗ ≤ q. (Here tr denotes thenormalised trace on Mn, i.e. tr((aij)ij) = 1

n

∑ni=1 aii.)

(c) Projections p and q are Murray–von Neumann equivalent if there is a partialisometry v ∈ A with v∗v = p and vv∗ = q. Check that this is an equivalencerelation. If A = Mn describe the equivalence classes.

STRUNG – INVITATION TO C∗-ALGEBRAS 29

4.2 Let A be a separable C∗-algebra. For two positive elements a, b in A we saythat a is Cuntz subequivalent b and write a . b if there are (vn)n∈N ⊂ A such thatlimn→∞ ‖vnbv∗n − a‖ = 0. We write a ∼ b and say a and b are Cuntz equivalent ifa . b and b . a.

(a) Show that ∼ is an equivalence relation on the positive elements.

(b) Show that if p and q are projections then they are Cuntz equivalent if theyare Murray–von Neumann equivalent.

(c) Let f, g ∈ C(X)+ where X is a locally compact metric space. Show that ifsupp(f) ⊂ supp(g) then for any ε > 0 there is a positive function e ∈ C(X) suchthat ‖f − ege‖ < ε. Now show that f . g if and only if supp(f) ⊂ supp(g).

(d) Let a ∈ A+ where a is a separable C∗-algebra. Show that a . an for everyn ∈ N. Use this to show that a∗a ∼ aa∗ for every a ∈ A.

4.3 Let A and B be concrete C∗-algebras. A linear map φ : A → B is positive ifφ(A+) ⊂ B+. For any map φ : A → B and n ∈ N we can define φ(n) : Mn(A) →Mn(B) by applying φ entry-wise. If φ(n) : Mn(A) → Mn(B) is positive for everyn ∈ N then we say φ is completely positive.

(a) Show that a ∗-homomorphism φ : A→ B is completely positive.

(b) Let A = B = M2 and let τ : M2 → M2 be the map taking a matrix to itstranspose. Show that τ is positive but not completely positive. (Note that weneed not restrict ourselves to concrete C∗-algebras, but we haven’t yet shown thatMn(A) is actually a C∗-algebra when we only know A is abstract.)

(c) Let φ : A→ B be a ∗-homomorphism and let v ∈ B. Show that the map

v∗φv : A→ B : a 7→ v∗φ(a)v

is completely positive.

4.4 Let X be a compact metric space. Show that for any finite subset F ⊂C(X) and any ε > 0 there is a finite dimensional C∗-algebra F and completelypositive contractive (c.p.c.) maps ψ : C(X) → F and φ : F → C(X) such that‖φ ψ(f)− f‖ < ε for every f ∈ F .

4.5 Suppose d ∈ N and thatX has covering dimension d, that is, every open cover Uof X has a finite subcover U ′ such that

∑U∈U ′ χU(x) ≤ d+1 for every x ∈ X. (Here

χU is the indicator function on U .) Show that for every finite subset F ⊂ C(X)and every ε > 0 there are finite dimensional C∗-algebras F0, . . . Fd and c.p.c. mapsψ : C(X)→

⊕di=0 Fi and φi : Fi → C(X) such that ‖(⊕di=0φi) ψ(f)− f‖ < ε for

every f ∈ F and moreover that φi(f)φi(g) = 0 whenever f 6= g.

4.6 Prove that every separable C∗-algebra admits a countable approximate unit.

4.7 Show that if a ∈ A+ then aAa is a hereditary C∗-subalgebra.

STRUNG – INVITATION TO C∗-ALGEBRAS 30

4.8 Let A be a unital C∗-algebra, a ∈ Inv(A) and p ∈ A a projection. Show thatif a commutes with p then a is invertible in the corner pAp. If a ∈ A is invertiblein pAp, is a ∈ Inv(A)?

5. Positive linear functionals and representations

5.1 A linear map φ : A→ B between C∗-algebras is called positive if φ(A+) ⊂ B+.Note that a ∗-homomorphism is always positive. If φ : A → C is positive, it iscalled a positive linear functional. Any positive linear functional φ also satisifiesφ(Asa) ⊂ R and hence φ(a∗) = φ(a) for every a ∈ A.

If φ is a positive linear functional with ‖φ‖ = 1, then φ is called a state and if itsatisfies φ(ab) = φ(ba) for every a, b ∈ A then it is called a tracial state. The setof states, respectively tracial states, on A will be denoted by S(A), respectivelyT (A).

5.2 Example: Let A = Mn. Then the usual normalised trace, tr : A→ C givenby

tr((aij)ij) =1

n

n∑i=1

aii

is a tracial state.

5.3 Example: Let A = C(X) be a commutative C∗-algebra. Then any characteris a tracial state, but not every tracial state is of this form. Suppose that µ is aprobability measure on X. Then the map τ : A→ C given by

τ(f) =

∫fdµ

is a tracial state. One often thinks of a tracial state as a noncommutative measure.

5.4 Example: Let H be a Hilbert space and ξ ∈ H a nonzero vector. Then

φ(a) = 〈aξ, ξ〉

is a postive linear functional, but not necessarily tracial.

The next theorem is another example of how C∗-algebras are in general morewell-behaved than arbitrary Banach algebras.

5.5 Theorem: Any positive linear functional on a C∗-algebra is bounded.

Proof: Suppose not. Let φ : A → C be an unbounded linear functional onsome C∗-algebra A. Since φ is unbounded, we can find a sequence (an)n∈N ofelements in the unit ball of A such that |φ(an)| → ∞ as n → ∞. Withoutloss of generality, we may assume that each an ∈ A+, for if φ was bounded onevery an ∈ A+ then φ would be bounded everywhere. Passing to a subsequence if

STRUNG – INVITATION TO C∗-ALGEBRAS 31

necessary, we may further assume that for every n ∈ N we have φ(an) ≥ 2n. Leta =

∑n∈N 2−nan ∈ A+. Then, for every N ∈ N,

φ(a) =∑n∈N

2−nφ(an) ≥∑n∈N

1 > N,

which is impossible.

Positive linear functionals admit the following Cauchy–Schwarz inequality.

5.6 Proposition: Let φ : A → C be a positive linear functional on the C∗-algebra A. Then

|φ(a∗b)|2 ≤ φ(a∗a)φ(b∗b)

for every a, b ∈ A.

Proof: We may obviously assume that φ(a∗b) 6= 0. Since φ is positive, for anyλ ∈ C we have φ((λa+ b)∗(λa+ b)) ≥ 0. In particular, this holds for

λ = t|φ(a∗b)|φ(b∗a)

,

for any t ∈ R, giving

t2φ(a∗a) + 2t|φ(a∗b)|+ φ(b∗b) ≥ 0.

If we have φ(a∗a) = 0 then φ(b∗b) ≥ 2t|φ(a∗b)| for every t ∈ R, which is impossibleunless |φ(a∗b)| is also zero; in this case the inequality holds. If φ(a∗a) 6= 0 then let

t = −|φ(a∗b)|φ(a∗a)

.

Then|φ(a∗b)|2

φ(a∗a)− 2|φ(a∗b)|2

φ(a∗a)+ φ(b∗b) ≥ 0,

from which the result follows.

5.7 Proposition: Let A be a C∗-algebra and let (uλ)λ∈Λ be an approximate unit.Let φ ∈ A∗. Then φ is positive if and only if limλ φ(uλ) = ‖φ‖. In particular if Ais unital then φ(1) = ‖φ‖.Proof: Suppose that φ is positive. Then, since (uλ)Λ is an increasing net ofpositive elements, (φ(uλ))Λ is increasing in R+. Since it’s moreover bounded, itconverges to some r ∈ R+. Since each uλ has norm less than one, r ≤ ‖φ‖. Nowif a ≥ 0 with ‖a‖ ≤ 1, then, using the Cauchy–Schwarz inequality,

|φ(auλ)|2 ≤ φ(a∗a)φ(u2λ) ≤ φ(a∗a)φ(uλ) ≤ rφ(a∗a).

Since φ is continuous and auλ → a, it follows that

|φ(a)|2 ≤ rφ(a∗a) ≤ r‖φ‖,hence ‖φ‖2 ≤ r‖φ‖ which is to say limλ φ(uλ) = ‖φ‖.

STRUNG – INVITATION TO C∗-ALGEBRAS 32

For the converse, first let a ∈ Asa. We will show that φ(a) ∈ R. Let α and βbe real numbers such that φ(a) = α+ iβ. Without loss of generality, assume thatβ ≥ 0. For n ∈ N let λ be sufficiently large that ‖uλa− auλ‖ < 1/n. Then

‖nuλ − ia‖2 = ‖nu2λ − a2‖

= ‖n2u2λ + a2 − in(uλa− auλ)‖

≤ n2 + 2.

We also have that

limλ|φ(nuλ − ia)|2 = (n‖φ‖+ β − iα)(n‖φ‖+ β + iα)

= (n‖φ‖+ β)2 + α2.

Thus

(n‖φ‖+ β)2 + α2 = limλ|φ(nuλ − ia)|2 ≤ ‖φ‖2‖nuλ − ia‖2

≤ ‖φ‖2(n2 + 2),

and then

n2‖φ‖2 + 2n‖φ‖β + β2 + α2 ≤ n2‖φ‖2 + 2,

for every n ∈ N. Since β2 + α2 ≥ 0 and we assumed that β ≥ 0, for large enoughn this can only hold with β = 0. Thus φ(a) ∈ Asa.

Now if a ∈ A+ and ‖a‖ ≤ 1, then uλ−a ∈ Asa and uλ−a ≤ uλ so limλ φ(uλ−a) ≤‖φ‖. Thus φ(a) ≥ 0.

5.8 Corollary: Let A be a nonunital C∗-algebra. Then any positive linearfunctional φ : A→ C admits a unique extension φ : A→ C with ‖φ‖ = ‖φ‖.Proof: Exercise.

5.9 Proposition: Let A be a C∗-algebra and let B ⊂ A a C∗-subalgebra. Everypositive linear functional φ : B → C admits an extension φ : A → C. If B ⊂A is a hereditary C∗-subalgebra then the extension is unique and if (uλ)Λ is anapproximate unit for B then

φ(a) = limλφ(uλauλ),

for any a ∈ A.

Proof: By the previous corollary, we may suppose that A is unital and 1A ∈ B.Then, applying the Hahn-Banach theorem there is an extension φ ∈ A∗ such that‖φ‖ = ‖φ‖. Since ‖φ‖ = φ(1) = φ(1), it follows from Proposition 5.7 that φ is alsopositive.

STRUNG – INVITATION TO C∗-ALGEBRAS 33

If B is hereditary, then limλ φ(uλ) = ‖φ‖ = ‖φ‖ = φ(1). Thus limλ φ(uλ−1) = 0.For any a ∈ A,

|φ(a)− φ(uλauλ)| ≤ |φ(a− auλ)|+ |φ(auλ − uλauλ)|≤ φ(a∗a)1/2φ((1− uλ)2)1/2 + φ(uλa

∗auλ)1/2φ((1− uλ)2)1/2

≤ φ((1− uλ))1/2(φ(a∗a)1/2 + φ(uλa∗auλ)

1/2)

= 0.

5.10 Proposition: Let A be a nonzero C∗-algebra and let a ∈ A be a normalelement. Then there is a state φ ∈ S(A) such that φ(a) = ‖a‖.Proof: Let B ⊂ A be the C∗-subalgebra generated by a and 1. Since B iscommutative, we have a ∈ C(Ω(B)) and φ ∈ Ω(B) such that φ(a) = a(φ) = ‖a‖.Since φ(1) = 1, there is a positive extension to A. Then the restriction φ to Asatisfies the requirements, since ‖φ‖ = φ(1) = 1.

5.11 A representation of a C∗-algebra A is a (H, π) consisting of a Hilbert spaceH and a ∗-homomorphism π : A → B(H). If π is injective then we say (H, π)is faithful. If A has a faithful representation, then it is ∗-isomorphic to a closedself-adjoint subalgebra of B(H), that is, a concrete C∗-algebra.

5.12 We begin by establishing that every C∗-algebra has many representations.Let A be a C∗-algebra. Given a positive linear functional φ, let

Nφ := a ∈ A | φ(a∗a) = 0,

which is a closed left ideal in A.

Define a map 〈·, ·, 〉φ : A/Nφ × A/Nφ → C→ C by 〈a+Nφ, b+Nφ〉 → φ(b∗a).

Exercise: Show that 〈·, ·, 〉φ defines an inner product on A/Nφ.

5.13 Let Hφ denote Hilbert space obtained by completing A/Nφ with respect tothe inner product described above. For a ∈ A, we define a linear operator πφ(a) :A → B(A/Nφ) by πφ(a)(b) = ab + Nφ. It is a straightforward exercise to checkthat this is a bounded ∗-homomorphism and hence extends to a ∗-homomorphism

πφ : A→ B(Hφ).

5.14 Definition: Let φ be a positive linear functional on the C∗-algebra A.The representation (Hφ, πφ) is called the Gelfand–Naimark–Segal representationassociated to φ, or more commonly the GNS representation associated to φ.

5.15 Let (Hλ, πλ)λ∈Λ be a family of representations of the C∗-algebra A. Define

⊕λ∈Λπλ : A→ ⊕λ∈ΛHλ

STRUNG – INVITATION TO C∗-ALGEBRAS 34

to be the map taking a ∈ A to the element with πλ(a) in the λ coordinate. Then(⊕λ∈ΛHλ,⊕λ∈Λπλ) is a representation of A. It is faithful as long as, for eacha ∈ A \ 0, there is some λ such that πλ(a) 6= 0.

5.16 Definition: Let A be a C∗-algebra. The representation

(⊕φ∈S(A)Hφ,⊕φ∈S(A)πφ)

is called the universal representation of A.

5.17 From the GNS construction above, we get the following Gelfand–Naimarktheorem.

Theorem: Let A be a C∗-algebra. Then its universal representation is faithful.

Proof: Exercise.

5.18 If A ⊂ B(H) is a C∗-subalgebra, then its commutant is defined as

A′ := b ∈ B(H) | ab = ba for every a ∈ A.

Clearly we have that A ⊂ A′′. If in fact A contains 1B(H) and A = A′′, then A iscalled a von Neumann algebra. Since (A′′)′′ = A′′ for any C∗-algebra A ⊂ B(H),A′′ is itself a von Neumann algebra. We call A′′ the enveloping von Neumannalgebra of A, or the double commutant of A.

5.19 Since every C∗-algebra is A contained in a von Neumann algebra A′′, we maythink of von Neumann algebras as “bigger” in a certain sense. That A ⊂ A′′ israther algebraic in nature, but this can also be seen topologically. In additionto the norm topology on B(H), we also have two weaker topologies: the weakoperator and strong operator topologies.

Definition: The weak operator topology on B(H) is the weakest topology suchthat the sets W (a, ξ, µ) = b ∈ B(H) | |〈aξ, µ〉| < 1 are open. The strongoperator topology is the weakest topology in which the sets S(a, ξ) = b ∈ B(H) |‖(a− b)ξ‖ < 1 are open.

The sets W (ai, ξi, µi | 1 ≤ i ≤ n) = ∩n1W (ai, ξi, µi) form a base for the weakoperator topology and the similarly, sets of the form W (ai, ξi | 1 ≤ i ≤ n) =∩ni S(ai, ξi) are a base for the strong operator topology.

5.20 Proposition: Let (aλ)λ∈Λ be a net in B(H) and let a ∈ B(H). Then

(i) (aλ)Λ converges to a in the weak operator topology, written aλWOT−−−→ a, if

limλ〈aλξ, η〉 = 〈aξ, η〉 for all ξ, η ∈ H;

(ii) (aλ)Λ converges to a in the strong operator topology, written aλSOT−−→ a, if

limλ aλξ = aξ for all ξ ∈ H.

Proof: Exercise.

STRUNG – INVITATION TO C∗-ALGEBRAS 35

5.21 Exercise: Show that convergence in the norm operator topology impliesconvergence in the strong topology implies convergence in the weak topology. Showthat the reverse implications do not neccessarily hold.

5.22 Proposition: Let a ∈ B(H). Left and right multiplication by a is bothSOT- and WOT-continuous.

Proof: Suppose that (bλ)Λ is a net in B(H) which WOT-converges to b. Letξ, µ ∈ H. Then

limλ〈abλξ, µ〉 = lim

λ〈bλξ, a∗µ〉

= 〈bξ, a∗µ〉= 〈abξ, µ〉,

showing abλWOT−−−→ ab. Thus multiplication on the left is WOT-continuous. The

other calculations are similar and are left as exercises.

5.23 Proposition: Let S be a subset of B(H). Then the commutant S ′ = a ∈B(H) | ax = xa for all x ∈ S is closed in the weak operator topology.

5.24 Definition: If A ⊂ B(H) then we say that A acts nondegenerately on Hif its null space

NA := ξ ∈ H | aξ = 0 for every a ∈ A

is trivial.

5.25 Now we can show von Neumann’s Double Commutant Theorem.

Theorem: Let A be a C∗-algebra acting nondegenerately on H. The followingare equivalent.

(i) A is a von Neumann algebra;

(ii) AWOT

= A;

(iii) ASOT

= A.

Proof: We have that ASOT ⊂ A

WOTsince the strong operator topology is stronger

than the weak topology. Since A ⊂ A′′ and A′′ is WOT-closed, we furthermore

have AWOT ⊂ A′′. It only remains to show that A′′ ⊂ A

SOT. Let a ∈ A′′. It is

enough to show that for any n ∈ N \ 0 and ξ1, . . . , ξn ∈ H there is some b ∈ Asuch that b is contained in S(a, ξi | 1 ≤ i ≤ n). Notice that this is equivalent tofinding b ∈ A with

∑ni=1 ‖(b− a)ξi‖2 < 1.

Consider n = 1. Let p be the orthogonal projection onto Aξ1. Then if c ∈ A wehave pcpξ1 = cpξ1 and if µ 6= ξ then pcpµ = cpµ = 0. Thus pcp = pc for everyc ∈ A. Thus cp = (pc∗)∗ = (pc∗p)∗ = pcp = pc whence p ∈ A′. Note that thismeans p⊥ ∈ A′, too.

STRUNG – INVITATION TO C∗-ALGEBRAS 36

If µ = p⊥ξ1 then Aµ = Ap⊥ξ1 = p⊥Aξ1 = 0. Since A acts nondegenerately, wemust therefore have µ = 0. It follows that ξ1 ∈ Aξ1. Since a ∈ A′′ we have thatpa = ap and aξ1 ∈ Aξ1. Thus we can find b ∈ A which satisfies ‖(a− b)ξ1‖ < 1.

Now suppose n ≥ 2. Let H(n) := H⊕· · ·⊕H be the direct sum of n copies of H.An arbitrary operator in B(H(n)) then looks like an n×n matrix (xij)ij with eachentry xij ∈ B(H). For c ∈ A, let c(n) ∈ B(H(n)) be defined as c(n)(ξ1, . . . , ξn) :=(cξ1, . . . , cξn) and then set A(n) := c(n) | c ∈ A.

We claim that (A(n))′′ = (A′′)(n) where (A′′)(n) is defined analogously to theabove. It is easy to see that c = (cij)ij ∈ (A(n))′ if and only if each cij ∈ A′.Thus A(n) contains all the matrix units eij where eij is the matrix with 1 in the(i, j)-entry and zero elsewhere. It follows that any c ∈ (A(n))′′ must commute withevery eij. The only way this is possible if all the diagonal entries of c are the

same and the off-diagonal entries are zero, that is c = c(n)11 where c11 ∈ A′′. It is

clear that c commutes with each b(n) where b ∈ A′. Thus c11 ∈ A′′ and c ∈ (A′′)n,proving the claim.

Now we apply the case for n = 1 to a(n) ∈ (An)′′ and ξ = (ξ1, . . . , ξn) to findb ∈ A with

1 > ‖(a(n) − b(n))ξ‖2 =n∑i=1

‖(a− b)ξi‖2.

Thus b is in S(a, ξi | 1 ≤ i ≤ n).

5.26 Proposition: The weak operator continuous linear functionals and thestrong operator continuous linear functionals φ : B(H) → C coincide and arealways of the form

φ(a) =n∑i=1

〈aξi, ηi〉

for some n ∈ N and ξi, ηi ∈ H, 1 ≤ i ≤ n.

Proof: It is easy to see that anything of the form φ(a) =∑n

i=1〈aξi, µi〉 will be aWOT and hence SOT continuous linear functional. Suppose now that φ is a SOT-continuous positive linear functional. Then the set U := a ∈ B(H) | |φ(a)| < 1is open in the strong operator topology and contains zero. Thus we can find abasic set as given in Definition 5.19 containing zero that is completely containedin U . That is to say, there are ξ1, . . . , ξn ∈ H such that

V := a ∈ B(H) | ‖aξi‖ < 1, 1 ≤ i ≤ n ⊂ a ∈ B(H) | |φ(a)| < 1.

If a ∈ B(H) satisfies∑N

i=1 ‖aξi‖ < 1 then clearly a ∈ V . Let H(n) = H ⊕ · · · ⊕Hbe the Hilbert space that is given by the direct sum of n copies of H. Set ξ :=(ξ1, . . . , ξn) and let ψξ : B(H)→ H(n) be the map given by ψξ(a) = (aξ1, . . . , aξn).Now define

F : ψξ(B(H))→ C

STRUNG – INVITATION TO C∗-ALGEBRAS 37

by F (ψξ(a)) = φ(a). Note that if η ∈ ψξ(B(H)), then η = (aξ1, . . . , aξn) for somea ∈ A and thus if ‖η‖ ≤ 1 we have

∑ni=1 ‖aξi‖ ≤ 1. In that case φ(a) ≤ 1. Thus

‖F‖ ≤ 1 and, by applying the Hahn–Banach theorem, there is an extension to acontinuous linear functional F : H(n) → C. By the Riesz representation theorem,there is η = (η1, . . . , ηn) ∈ H(n) such that

F (ν) = 〈ν, η〉H(n) =n∑i=1

〈νi, ηi〉H ,

and so

φ(a) = F (ψ(a)) =n∑i=1

〈aξi, ηi〉,

and the result follows since this is also WOT continuous.

5.27 Proposition: Let A ⊂ B(H) be a C∗-algebra. Suppoes that K ⊂ H is anA-invariant subspace, that is AK = aξ | a ∈ A, ξ ∈ K ⊂ K. Let p ∈ B(H) bethe orthogonal projection onto K. Then p ∈ A′.Proof: Since K is invariant apξ ∈ K for every a ∈ A and every ξ ∈ H. Thuspapξ = apξ for every a ∈ A and ξ ∈ H, so pa = (a∗p)∗ = (pa∗p)∗ = pap = ap.Hence p ∈ A′.5.28 Lemma: Let A ⊂ B(H) be a C∗-algebra and I ⊂ A an ideal. Then theprojection p ∈ B(H) onto the closed subspace IH is in A′.

Proof: Since A(IH) = (AI)H = AH, the subspace IH is invariant. The restfollows from the previous proposition.

5.29 The following is a reformulation of the Hahn–Banach separation theoremwhich we will require in the proof of Theorem 4.15.

Lemma: Let V be a locally compact topological vector space and let S ⊂ V be aclosed convex subset. Then if x /∈ S the sets x and S are strictly separated, thatis, there is a continuous linear functional φ ∈ V ∗, r ∈ R and ε > 0 such that

<(φ(x)) < r < r + ε < <(φ(s)) for every s ∈ S.

5.30 We are now able to prove Theorem 4.15: that every ideal I in a C∗-algebraA has an approximate unit which is quasicentral for A.

Proof of Theorem 4.15: Let A be a C∗-algebra, I ⊂ A an ideal and letπ : A → B(H) be the universal representation of A. Since (H, π) is faithful, it isenough to show the theorem holds for π(A) and π(I).

Let (uλ)Λ be an approximate unit for π(I) ⊂ π(A). Let E denote the convex hullof uλ | λ ∈ Λ, that is,

E = ∑n

i=1 µixi |∑n

i=1 µi = 1 and xi = uλ for some λ ∈ Λ.

STRUNG – INVITATION TO C∗-ALGEBRAS 38

If∑n

i=1 µixi ∈ E then there is some uλ such that xi ≤ uλ for every 1 ≤ i ≤ n.Thus

∑ni=1 µixi ≤

∑ni=1 µiuλ = uλ, which shows that E is upwards directed and is

furthermore an approximate unit for I.

We will show that for every λ0 ∈ Λ and finite set of elements a1, . . . , an ∈ π(A)there is e ∈ E e ≥ uλ0 with

‖aie− eai‖ < 1/n for every 1 ≤ i ≤ n.

So let a1, . . . , an and uλ0 be given. Let H(n) denote the Hilbert space givenby the direct sum of n copies of H and for an element a ∈ π(A), denote bya(n) ∈ Mn(π(A)) the element with a copied n times down the diagonal. Notice

that (u(n)λ )Λ is an approximate unit for Mn(π(I)). Let Eλ≥λ0 denote the convex

hull of uλ | λ ≥ λ0. Let c = a1 ⊕ · · · ⊕ an.

Set

S := ce(n) − e(n)c | e ∈ Eλ≥λ0.

We claim that 0 ∈ S. To show this, suppose it is not. Since Eλ≥λ0 is convex,it is easy to see that S is convex. Then by the Lemma 5.29 there is a continuouslinear functional φ ∈ (B(H))∗, r ∈ R and ε > 0 such that

0 = <(φ(0)) < r < r + ε < <(φ(s))

for every s ∈ S. Rescaling if necessary, we may assume that 1 ≤ <(φ(s)) for everys ∈ S. Thus there is ξ, η ∈ H(n) such that φ(a) = 〈aξ, η〉.

We have uλSOT−−→ p where p is the projection onto π(I)H. Since p ∈ π(A)′,

φ(cu(n)λ − u

(n)λ c) = 〈(cu(n)

λ − u(n)λ c)ξ, η〉

→ 〈(cp(n) − p(n)c)ξ, η〉= 0,

contradicting the fact that φ(s) ≥ 1 for every s ∈ S. Thus we must have that0 ∈ S, which proves the claim.

It follows that for any n ∈ N \ 0, finite subset F = a1, . . . , an and λ0 ∈ Λthere is fF ,λ0 ∈ Eλ≥λ0 such that

‖(a1 ⊕ · · · ⊕ an)f(n)F ,λ0 − f

(n)F ,λ0(a1 ⊕ · · · ⊕ an)‖ < 1/n.

Thus

‖aifF ,λ0 − fF ,λ0ai‖ ≤ max1≤j≤n

‖aifF ,λ0 − fF ,λ0ai‖

= ‖(a1 ⊕ · · · ⊕ an)f(n)F ,λ0 − f

(n)F ,λ0(a1 ⊕ · · · ⊕ an)‖

< 1/n.

STRUNG – INVITATION TO C∗-ALGEBRAS 39

It follows that (fF ,λ)F ,λ, where F runs over all finite subsets of π(A) and λ ∈ Λ,is a quasicentral unit approximate unit for π(I).

Exercises

5.1 Show that any positive linear functional φ : A→ C admits a unique extensionφ : A→ C such that ‖φ‖ = ‖φ‖.5.2 Let A = Mn and H = Cm for n,m ∈ N. For what values of m does A admita faithful representation on H? Show that if π1, π2 : Mn → Mm are both faithfulrepresentations on Cm, then they are unitarily equivalent: there exists a unitaryu ∈Mm such that π1(a) = u∗π2(1)u for every a ∈Mn.

5.3 Prove that the universal representation of a C∗-algebra A is always faithful.Thus abstract C∗-algebras and concrete C∗-algebras coincide.

5.4 Show that for any C∗-algebra A, Mn(A) is also a C∗-algebra. Show thatMn(A) ∼= Mn⊗A, where ⊗ denotes the algebraic tensor product. (See for example[2].) In general, there is more than one way to take the tensor product of C∗-algebras in such a way that their algebraic tensor product is a dense ∗-subalgebra.If A is a C∗-algebra such that A ⊗1 B ∼= A ⊗2 B for any C∗-algebra B and anyC∗-tensor product is called nuclear. Deduce that Mn is a nuclear C∗-algebra.

5.5 Let H be a Hilbert space. In B(H), show that convergence in the operatornorm topology implies strong operator convergence which in turn implies weakoperator convergence.

5.6 Show that the commutant S ′ of a subset S ⊂ B(H) is closed in the weakoperator topology. Show that if S = S∗ then S ′ is a ∗-algebra.

5.7 Let A be a C∗-algebra and let (H, π) be the universal representation of A.

Show that π(A)SOT

is a von Neumann algebra.

5.8 Let A ⊂ B(H) be C∗-algebra with approximate unit (uλ)Λ. Does (uλ)Λ con-verge in the strong operator or weak operator topology?

5.9 Let (aλ)Λ ⊂ B(H) be a net that is WOT-convergent. Show that (aλ)Λ mustbe norm bounded.

5.10 Let B be a strongly closed hereditary subalgebra of A. Show that there is aunique projection p ∈ B such that B = pAp.

6. Further examples of C∗-algebras

UHF algebras and AF algebras. 6.1 A C∗-seminorm on a ∗-algebra A is aseminorm p on A such that, for all a and b in A, we have p(ab) ≤ p(a)p(b),p(a∗) = p(a) and p(a∗a) = p(a)2.

STRUNG – INVITATION TO C∗-ALGEBRAS 40

Let p : A → R+ be a C∗-seminorm on a ∗-algebra A. Then N = ker(p) is aself-adjoint ideal in A, and this induces a C∗-norm on the quotient A/N given by

‖a + N‖ = p(a). Let B = A/N‖·‖

be the completion with respect to this norm.Define the multiplication and involution in the obvious way. This makes B into aC∗-algebra called the enveloping C∗-algebra of (A, p).

The map i : A→ B : a→ a+N is called the canonical map and the image of Aunder i is a dense ∗-subalgebra of B.

6.2 An inductive sequence of C∗-algberas (An, φn)n∈N consists of a sequence ofC∗-algebras (An)n∈N and a sequence of connecting ∗-homomorphisms (φn : An →An+1)n∈N

6.3 Proposition: Let (An, φn)n∈N be an inductive sequence of C∗-algebras. Let

A = a = (aj)j∈N ⊂ Πj∈NAj | there is N ∈ N such that aj+1 = φj(aj) for all j ≥ N.Then A is a ∗-algebra under pointwise operations and

p : A → R+ : a 7→ limk→∞‖ak‖Ak

is a C∗-seminorm on A.

Proof: Exercise.

6.4 Definition: Let (An, φn)n∈N be an inductive sequence of C∗-algebras. Theinductive limit of (An, φn)n∈N, written lim−→(An, φn), (or simply lim−→An if it’s clear

what the maps should be) is the enveloping C∗-algebra of (A, p), where A and pare the ∗-algebra and C∗-seminorm, respectively, as defined in Proposition 6.3.

6.5 Let (An, φn)n∈N be an inductive sequence of C∗-algebras and let A = lim−→An be

the inductive limit. It is useful to describe maps between nonadjacent C∗-algebrasin the inductive sequence, as well as from each An in the sequence to the A. Thuswe define, for n < m

φn,m : An → Am

to be the composition

φm−1 · · · φn+1 φn, .If a ∈ An, then define (aj)j∈N ⊂ Πj∈NAj by

aj =

0 if j < n,a if j = n,

φn,j−1(a) if j > n

Clearly (aj)j∈N ∈ A. From this we define the map

φ(n) : An → A

by φ(n)(a) = ι((aj)j∈N where ι : A → A/Np is the canonical map from A into itsenveloping C∗-algebra A. From this we get, for every n,m n < m, a commutative

STRUNG – INVITATION TO C∗-ALGEBRAS 41

diagram

Anφn,m //

φ(n) !!

Am

φ(m)

A

This leads to the following universal property.

6.6 Theorem: Let (An, φn)n∈N be an inductive sequence of C∗-algebras with limitA = lim−→An. Suppose there is a C∗-algebra B and for every n ∈ N there are∗-hommoorphisms ψ(n) : An → B making the diagrams

Anφn //

ψ(n)""

An+1

ψ(n+1)

B

commute. Then there is a unique ∗-homomorphism ψ : A → B making the dia-grams

Anφ(n) //

ψ(n)

A

ψB

commute.

Proof: Let B and the ∗-homomorphisms ψ(n) : An → B be given. If (aj)j∈N ∈ A,then there is N ∈ N such that for every j ≥ N we have aj+1 = φj(aj). Bycommutativity of the first diagram we have that ψN(aN) = ψj(aj) for every j ≥ N .

Suppose that a ∈ An, b ∈ Am and φ(n)(a) = φ(m)(b) ∈ lim−→An. If n ≤ m then

φ(m) φn,m(a) = φ(m)(b) by commutativitiy of the diagram in (6.5). It follows thatlimk→∞,k≥m ‖φn,k(a)−φm,k(b)‖ = 0. Thus limk→∞ ‖ψk(φn,k(a))−ψk(φm,k(b))‖ = 0.

The above shows that ψ : ι(A) → B defined on each (aj)j∈N by ι ψN(aN) forsufficiently large N , is well-defined and extends to a ∗-homomorphism ψ : lim−→A→B, making the second diagram commute.

6.7 Exercise: Let A = lim−→(An, φn) be the inductive limit of a sequence of C∗-

algebras and let B be a C∗-algebra. For every n ∈ N let ψ(n) : An → B be a∗-homomorphism satisfying ψ(n+1) φn = ψ(n) and let ψ : A → B be the induced∗-homomorphism. Then

(i) ψ is injective if and only if ker(ψ(n)) ⊂ ker(φ(n)) for every n ∈ N, and

(ii) ψ is surjective if and only if B = ∪∞j=1ψ(n)(An).

STRUNG – INVITATION TO C∗-ALGEBRAS 42

6.8 A supernatural number p is given by the infinite product p = Πp prime pkp

where kp ∈ N ∪ ∞. Every natural number is thus a supernatural number. Asupernatural number is of infinite type if, for every prime p, we have either kp = 0or kp =∞.

6.9 To every supernatural number p, we may associate an inductive system ofmatrix algebras as follows. We may choose natural numbers (ni)i∈N such that, foreach i, ni divides ni+1 and ni divides p, but ni does not divide pp for any primep unless p∞ divides p. Call such a sequence a UHF decomposition of p. For eachni, let φi : Mni → Mni+1

be the map that sends a matrix a of size ni to the blockmatrix of size ni+1 by copying a ni+1/ni times down the diagonal:

φ : Mni →Mni+1: a 7→

a 0 · · · 00 a · · · 0...

. . ....

0 0 · · · a

ni+1

nitimes.

6.10 Definition: Let p be a supernatural number and (nk)k ⊂ N a UHFsequence for p. The uniformly hyperfinite (UHF) algebra of type p is the inductivelimit of the inductive system (Mni , φi),

Up = lim−→Mni .

6.11 Exercise: Show that the above is well-defined.

6.12 Theorem: Let (An, φn) be an inductive limit of simple C∗-algebras. Thenlim−→An is simple.

Proof: A C∗-algebra A is simple if and only whenever B is another C∗-algebraand ψ : A → B is a surjective ∗-homomorphism, then φ is injective. Supposethen that ψ : lim−→An → B is a surjection onto a nonzero C∗-algebra B. For any

n ∈ N, φ(n)(An) ⊂ A is the image of a simple C∗-algebra and so is also simple.

Thus ψ|φ(n)(An) : φ(n)(An) → B is either zero or injective. Since A ∼= ∪n∈Nφ(n)(A)

and ∪n∈Nφ(n)(A) is dense, there must be some N ∈ N and a ∈ φ(N)(A) such thatψ(a) 6= 0. In this case ψ|φ(N)(AN ) must be injective.

Now, if k < N then φ(k)(Ak) ⊂ φ(n)(An). Thus ψ|φ(k)(Ak) is nonzero and hence

injective. If k > N then φN(AN) ⊂ φk(Ak). Then a ∈ φk(Ak) so again ψ|φ(k)(Ak)

is nonzero and hence injective. Thus ψ : A → B is injective on a dense subset,hence injective.

6.13 Corollary: A UHF algebra is simple.

STRUNG – INVITATION TO C∗-ALGEBRAS 43

6.14 An approximately finite-dimensional (AF) algebra generalises UHF algebras.Now in the inductive sequence, we allow any finite C∗-subalgebra F to appear, notonly matrix algebras.

Definition: An approximately finite-dimensional (AF) algebra is the inductivelimit of a sequence (Fn, φn) where Fn is a finite dimensional C∗-algebra for everyn ∈ N.

6.15 Note that a UHF algebra is an AF algebra, but the opposite need not bethe case. For example, the compact operators K is an AF algebra but is noteven unital. Still, AF algebras are a very nice class of C∗-algebras and are quitetractable: we can say a lot about their structure. We saw that a UHF algebra isuniquely determined by its associated supernatural number. For AF algebras, thesituation is slightly more complicated. Instead of a single supernatural number, werequire a so-called dimension group to distinguish AF algebras. In the simple unitalcase, this boils down to the pointed, ordered K0-group: the Grothendieck group ofMurray–von Neumann equivalence classes of projections in matrix algebras overthe AF algebra, together with the order on K0 induced by the order on positiveelements, as well as the class of the unit. Since K0 respects inductive limits, thisis an easily computable invariant for AF algebras. We’ll have more to say on thislater on, in the meantime, we have another nice structural property of AF algebras.

6.16 Recall that a partial isometry v in a C∗-algebra A is an element such thatboth v∗v and vv∗ are projections.

Definition: Let A be a C∗-algebra. If p, q ∈ A are projections, then p is Murray–von Neumann equivalent to q if there is a partial isometry v ∈ A such that v∗v = pand vv∗ = q. The projection p is Murray–von Neumann subequivalent to q if thereis a partial isometry v ∈ A such that v∗v = p and vv∗ ≤ q.

A projection p ∈ A is called finite if p is not Murray–von Neumann equivalentto a proper subprojection of itself, that is, there is no v ∈ A with v∗v = p andvv∗ ≤ p but vv∗ 6= p; otherwise p is said to be infinite.

A unital C∗-algebra A is called finite if 1A is finite. A unital C∗-algebra A iscalled stably finite if Mn(A) is finite for every n ∈ N.

6.17 An isometry in a unital C∗-algebra A is an element s ∈ A with s∗s = 1A.Clearly any unitary in A is an isometry, but an isometry need not be a unitary ingeneral. (See exercises.) If, however, this is the case, then A is finite:

Proposition: Let A be a unital C∗-algebra. Suppose that every isometry in Ais a unitary. Then A is finite.

Proof: Suppose that there is p ∈ A with p ≤ 1A and v ∈ A such that v∗v = 1Aand vv∗ = p. But v∗v = 1A means that v is an isometry and hence unitary. Thuswe have p = vv∗ = 1A which shows that 1A is finite.

STRUNG – INVITATION TO C∗-ALGEBRAS 44

6.18 Theorem: If A is a unital AF algebra, then A is stably finite.

Proof: Exercise.

6.19 As has been regularly mentioned, because of the fact that the Gelfand trans-form is a ∗-isomorphism, C∗-algebras are often thought of “noncommutative” lo-cally compact Hausdorff spaces. In the topological setting, there are several notionsof the topological dimension of a space. A point should be zero dimensional, aninterval one-dimensional, an n-cube n-dimensional, and so forth.

In what follows, we will denote the indicator function of an open set U by χU ;thus

χU(x) =

1 if x ∈ U,0 if x /∈ U.

Definition: Let X be a locally compact Hausdorff space. We say that X hascovering dimension d, written dim(X) = d if d is the least integer such that thefollowing holds: For every open cover O of X there is a finite refinement O′ suchthat, for every x ∈ X,

∑UO′ χU(x) ≤ d + 1. If no such d exists, we say that

dim(X) =∞.

In the case that we restrict ourselves to locally compact metrisable spaces, thevarious definitions coincide with the covering dimension.

Moving to the noncommutative setting, we would like to find an analogue ofthe dimension of a space. At the commutative level, metrisable corresponds tothe C∗-algebra being separable, so we content ourselves with trying to establishnoncommutative versions of covering dimension for separable C∗-algebras. Theseshould extend covering dimension in the sense that the noncommutative dimensionof C0(X) should be the same as the covering dimension. There are a few suchextensions; we describe three of them below.

6.20 Definition: Let A be a unital separable C∗-algebra A. We say that Ahas real rank d, written RR(A) = d, if d is the least integer such that, whenever0 ≤ n ≤ d + 1 the following holds: For every n-tuple (a1, . . . , an) of self-adjointelements in A and every ε > 0 there exists an n-tuple (y1, . . . , yn) ⊂ Asa such that∑n

k=1 y∗kyk is invertible and ‖

∑nk=1(xk − yk)∗(xk − yk)‖ < ε. If there is no such d,

then we say the real rank of A is infinite.

6.21 The stable rank has a very similar definition, dropping the fact that the n-tuples need be self-adjoint, and we only look at n-tuples from 0 ≤ n ≤ d (ratherthan d+ 1).

Definition: Let A be a unital separable C∗-algebra A. We say that A has stablerank d, writen SR(A) = d if d is the least integer such that, whenever 1 ≤ n ≤ dthe following holds: For every n-tuple (a1, . . . , an) of elements in A and everyε > 0 there exists an n-tuple (y1, . . . , yn) ⊂ A such that

∑nk=1 y

∗kyk is invertible

STRUNG – INVITATION TO C∗-ALGEBRAS 45

and ‖∑n

k=1(xk − yk)∗(xk − yk)‖ < ε. If there is no such d, then we say the realrank of A is infinite.

6.22 The nuclear dimension has a different flavour to the real and stable rank.It is a refinement of the completely positive approximation property, which, forC∗-algebras, is equivalent to nuclearity (that is, A has the completely positiveapproximation property if and only if A is nuclear).

Definition: Let A be a separable C∗-algebra. We say that A has nucleardimension d, written dimnucA = d, if d is the least integer satisiying the following:For every finite subset F ⊂ A and every ε > 0 there are a finite dimensional, C∗-algebra with d+ 1 ideals, F = F0 ⊕ Fd, and completely positive maps ψ : A→ Fand φ = F → A such that ψ is contractive, φ|Fn is contractive and orthogonalitypreserving (ie. for any a, b ∈ (Fn)+ with ab = ba = 0, we have φ(a)φ(b) = 0) and

‖φ ψ(a)− a‖ < ε for every a ∈ F .

If no such d exists, then we say dimnuc(A) =∞.A completely positive contractive map which is orthogonality preserving (such

as φ|Fn of the previous definition) is called order zero.

6.23 We should think of AF algebras of “zero-dimensional” objects. Note, however,that there is no definition for “stable rank zero”. We get the following theoremfor AF algebras. The proof is an exercise; see the exercise sheet for a bit of a hint.

Theorem: Let A be a unital AF algebra. Then A has real rank zero, stable rankone and nuclear dimension zero.

If A is nonunital, we can define its real and stable rank by putting RR(A) :=RR(A) and SR(A) := SR(A). Then the above theorem is also true for nonunitalAF algebras.

Group C∗-algebras. 6.24 Now that we know a bit about representations and thedifferent topologies on B(H) we can construct C∗-algebras out of locally compactgroups. We will also assume throughout this section that the groups are Haus-dorff. A topological group is a group together with a topology which makes groupoperations continuous. In particular, any group is a topological group with thediscrete topology.

6.25 A Borel measure on a group G is left-translation-invariant if, for any Borelset E ⊂ G and any s ∈ G, we have µG(sE) = µG(E).

Theorem: Let G be a locally compact group. There is a left-translation-invariantBorel measure on G, denoted µG, which is unique up to scalar multiple.

This measure is called a (left) Haar measure of G. When G is compact, µ(G) isfinite and so we can normalise it so that µ(G) = 1. If G is infinite and discrete,then we normalise so that µ(e) = 1.

STRUNG – INVITATION TO C∗-ALGEBRAS 46

6.26 In general, a left translation-invariant measure need not be right-translation-invariant. However, if this is the case, then we call G unimodular. Unimodulargroups include the cases that G is abelian, discrete, or compact. For the sake ofbrevity, we will stick to the unimodular case, though most of what we’ll do can begeneralised.

6.27 There are a number of noncommutative algebras that we can associate with agroup. The first is the group algebra of G, which is the algebra of formal C-linearcombinations and is denoted CG. The multiplication in CG extends the groupmultiplication.

6.28 We also have the function algebra of compactly supported functions on G,denoted Cc(G), which comes equipped with convolution as multiplication,

f ∗ g(t) =

∫G

f(s)g(s−1t)dµ(s),

and inversion for involution,

f ∗(s) = f ∗(s−1).

There is also a norm on Cc(G) given by ‖f‖1 =∫G|f(t)|dµ(t).

Notice that CG ⊂ Cc(G), but they are not the same unless G is discrete. WhenG is locally compact, Cc(G) will not be complete with respect to this norm. Com-pleting with respect to ‖ · ‖1 gives us yet another group algebra, L1(G, µ).

6.29 The space L1(G, µ) consists of functions f : G→ C such that∫G

|f(t)|dµ(t) <∞.

It is a Banach ∗-algebra when equipped with ‖ · ‖1, convolution and inversion asin the case of Cc(G).

6.30 We have CG ⊂ Cc(G) ⊂ L1(G, µ). Cc(G) is a dense ∗-subalgebra of L1(G).If G is finite, then L1(G, µ) ∼= CG. Since a left Haar measure is unique up to ascalar multiple, for any two left Haar measures µ, µ′ we have L1(G, µ) ∼= L1(G, µ′).Thus henceforth we will write L1(G) for any L1(G, µ) defined with respect to aleft Haar measure.

The ‖ · ‖1 norm is not a C∗-norm in general, so L1(G) is not a C∗-algebra. Thuswe would like to find a ∗-homomorphism from L1(G) into a C∗-algebra so that wecan complete the image of L1(G) to a C∗-algebra.

6.31 Proposition: Let G be a locally compact group. Then L1(G) is unital ifand only if G is discrete. In any case, L1(G) always has a norm one approximateunit.

Proof: If G is discrete, then a unit is given by the function that is 1 and theidentity e ∈ G and zero everywhere else. If L1(G) is unital then 1 ∗ f(e) =

STRUNG – INVITATION TO C∗-ALGEBRAS 47∫G

1(s)f(s−1)dµ(s) = f(e) for every f ∈ L1(G, µ) only if 1(s) = 0 for every s 6= e.But then if G is not discrete 1 = 0 a.e.µ, thus is not a unit in L1(G, µ).

For any G, let O be the collection of open neighbourhoods E of e (the identityin G). For E ∈ O, let fE be a function in L1(G) with f(e) = 1, supp(fE) ⊂ E,f ∗E = f ∗ and ‖f‖1 = 1. Since the set of all such neighbourhoods is upwards directedwith respect to reverse containment, (fE)E is an approximate unit for L1(G).

6.32 Definition: Let G be a locally compact group. A unitary representationof G is given by a pair (H, u) consisting of a Hilbert space H and a stronglycontinuous homomorphism u : G → U(H), where U(H) is the group of unitaryoperators on H. Here, strongly continuous means that g 7→ u(g)ξ is continuousfor every ξ ∈ H.

We say that a unitary representation is irreducible if u(G) does not commutewith any proper projections in B(H).

6.33 Let A ⊂ B(H) be a C∗-algebra. A is said to be irreducible whenever K ⊂ His closed vector subspace with AK ⊂ K, then K ∈ 0, H.Proposition: Let G be a locally compact group and u : G → U(H) a uni-tary representation. Then the C∗-subalgebra of B(H) generated by u(G), writtenC∗(u(G)), is irreducible if and only if u(G) is irreducible.

Proof: Exercise.

6.34 A representation (H, π) of A is called nondegenerate if the linear span ofπ(a)ξ | a ∈ A, ξ ∈ H, denoted by π(A)(H), is dense in H, or, equivalently, foreach ξ ∈ H \ 0 there is a ∈ A such that π(a)(ξ) 6= 0.

If u : G→ U(H) is a unitary representation then π : L1(G)→ B(H) given by

π(f)ξ =

∫G

f(t)ut(ξ)dµ(t), for f ∈ L1(G), ξ ∈ H

is a representation of L1(G). (Here we write ut to denote the operator u(t).)

Conversely, if we have a representation π : L1(G) → B(H) which is nondegen-erate then we can find a unique unitary representation of G as follows: Then wehave

limE∈O

π(fE)π(g)ξ = π(g)ξ

for every g ∈ L1(G). It follows that π(fE)SOT−−→ 1B(H).

Define u : G→ B(H) by

u(s)π(g)ξ = π(gs)ξ for s ∈ G, ξ ∈ H

where gs(t) = g(s−1t). In this case we have that u(s) = SOT limE∈O π((fE)s),which in turn implies that u is contractive. Furthermore, it is not hard to check

STRUNG – INVITATION TO C∗-ALGEBRAS 48

that u(s) is unitary for every s ∈ G. The construction ensures that this u isunique.

6.35 Definition: Let G be a locally compact group with Haar measure µ. Theleft regular representation of G on the Hilbert space L2(G), λ : G→ U(L2(G)), isgiven by

λ(s)f(t) = f(s−1t).

(Check that this is indeed a unitary representation of G.)

6.36 Definition: The reduced group C∗-algebra of G, written C∗r(G) is theclosure of λ(L1(G)) in B(L2(G)). The full group C∗-algebra, denoted C∗(G) isthe closure of L1(G) under the direct sum of all irreducible representations, or,equivalently, C∗(G) is the completion of L1(G) with respect to the norm

‖f‖ = sup‖π(f)‖ | π : L1(G, µ)→ B(H) is a ∗-representation.

6.37 Definition: Given a locally compact abelian group G, a character of G isa continuous group homomorphism from G → T. The set of all characters of Ghas the structure of a compact abelian group, which we call the Pontryagin dualof G and denote by G.

6.38 Let f ∈ L1(G). The Fourier–Plancheral transform f on G is given by

f(γ) =

∫G

γ(t)f(t)dµ(t).

Exercise: For f, g ∈ L1(G) we have (f ∗ g)ˆ = f g and (f ∗)ˆ = f .

6.39 We will require the following, which we use without proof in Theorem 6.40

Theorem: [Plancheral Theorem] The Fourier–Plancheral transform extends from

a map L1(G)→ L1(G) to a unitary operator from L2(G)→ L2(G).

6.40 Theorem: Let G be an abelian group. Then C∗(G) ∼= C∗r(G) ∼= C0(G).

Proof: Let f, g ∈ L1(G). Then f ∗ g(t) =∫Gf(s)g(s−1t)dµ(s) and putting

x = s−1t, we get s = tx−1 = x−1t so∫G

f(s)g(s−1t)dµ(s) =

∫G

f(x−1t)g(x)dµ(x) =

∫G

g(x)f(x−1t)dµ(x),

and we have f ∗ g = g ∗ f , that is, L1(G) is commutative.

Let Γ : L1(G)→ C0(Ω(L1(G)) be the Gelfand transform. Recall that Ω(L1(G))is the character space of L1(G) and notice that a character is exactly a one-dimensional representation. As we saw above, every representation of L1(G) cor-responds to a unitary representation of G on the same Hilbert space; here theHilbert space is C. The one-dimensional unitary representations of G are just the

STRUNG – INVITATION TO C∗-ALGEBRAS 49

characters of G, that is, G. Thus the Gelfand transform maps L1(G) → C0(G).Moreover, we have

f 7→ f

where f is the Fourier–Plancheral transform of f . The range of Γ(L1(G)) is clearlyself-adjoint. Moreover, is separates points by definition of the Fourier–Plancheraltransform. Thus Γ(L1(G)) is dense in C0(G). By Plancheral’s Theorem, this

extends to a unitary operator u : L2(G)→ L2(G). Then,

u(λ(f))u∗(g) = u(λ(f))g = (f ∗ g)ˆ = f g,

when f ∈ L1(G) and g ∈ L2(G) ∩ L1(G). Thus C0(G) 3 f → Mf ∈ B(L2(G)),

where Mf denotes the operator given by multiplication f . Since this is isometric,

λ is an isometric isomorphism. Thus C∗(G) ∼= C∗r(G).

Note that in this case, if G is discrete, then C∗r (G) can be given the additionalstructure of a compact quantum group. This is true more generally (for nonabeliandiscrete groups), and one of the particularly nice things about quantum groupsin general, is that their theory is able to extend this notion of Pontryagin dualitybeyond the case of abelian groups. Unfortunately, this is beyond the scope of thiscourse, but I think it is a nice motivation for the study of quantum groups.

Crossed products. An important an interesting generalisation of a group C∗-algebra is the crossed product of a C∗-algebra by a locally compact group G. Theconstruction of the crossed product has a lot of similarities to the construction ofgroup C∗-algebras, but now we have to take into account things like the represen-tations of the C∗-algebra. Again we will restrict ourselves to unimodular groups.

By an action α of G on a C∗-algebra A we always mean a strongly continuousgroup homomorphism α : G → Aut (A), where Aut (A) denotes the group of∗-automorphisms of A.

6.41 Definition: Suppose that G is a locally compact group acting on a C∗-algebra A. A covariant representation is a triple (H, π, u) where H is a Hilbertspace, (H, π) is a representation for A, (H, u) is a unitary representation for G andπ and u satisfy the covariance condition

u(g)π(a)u(g)∗ = π(αg(a)),

for every a ∈ A and g ∈ G.

6.42 For a group action α : G → Aut (A), the space L1(G,A, α) is defined asfollows.

First, take compactly supported continuous functions Cc(G,A) with twisted con-volution

(f ∗ g)(t) =

∫G

f(s)αs(g(s−1t)ds.

STRUNG – INVITATION TO C∗-ALGEBRAS 50

and

f ∗(s) = (αs(f(s−1)∗)

L1(G,A, α) is the completion with respect to the 1-norm

‖f‖1 =

∫G

‖f(s)‖ds.

Note that the norm inside the integral is the norm of the C∗-algebra A.

6.43 As we did for the group algebras, we can relate a covariant representation forα : G→ Aut (A) to a representation of L1(G,A, α) on an L2 space, which can bedefined as the completion of Cc(G,A, α) with respect to the norm

‖f‖2 :=

(∫G

‖f(s)‖2ds

)1/2

.

Given a covariant representation (H, π, u), the integrated form of (H, π, u) isdefined to be

π(f)(g) =

∫G

π(f(s))u(s)gds

where f ∈ L1(G,A, α) and g ∈ L2(G,A, α).

Proposition: Let G be a locally compact group, A a C∗-algebra and α : G →Aut (A) an action of G on A. For any covariant representation, the associatedintegrated form is a representation of L1(G,A, α) on H = L2(G,A, α).

In the opposite direction, a representation π : L1(G,A) also gives a covariantrepresentation of α : G→ Aut (A) using an approximate unit for L1(G,A). Sincethe details are similar to the case for unitary representations, they are left as anexercise.

6.44 Given a Hilbert space H0, the space L2(G,H0) consists of square-integrablefunctions from f to H0. It is a Hilbert space with respect to the inner product

〈f, g〉 =

∫G

〈f(s), g(s)〉H0ds.

The (left) regular covariant representation corresponding to π0 : A → H0 is(H, π, u) where H = L2(G,H0), π : A→ B(H) is given by

π(a)(f)(s) = π0(αs−1)f(s), (f ∈ L2(G,H0), s ∈ G);

and u : G→ U(H) is given by

u(s)(f)(t) := f(s−1t), (f ∈ L2(G,H0), t ∈ G).

6.45 Definition: Let α : G → Aut (A) be an action of locally compact groupG on a C∗-algebra. Let λ : L1(G,A, α) → B(H) denote the direct sum of all

STRUNG – INVITATION TO C∗-ALGEBRAS 51

integrated forms of regular representations. The reduced crossed product of A byG, written Aor,α G, is the closure of λ(L1(G,A, α)) ⊂ B(H).

The full crossed product of A by G is the closure of πu(L1(G,A, α)) ⊂ B(Hu)

where (πu, Hu) denotes the universal representation, that is, the direct sum of allirreductible representations of L1(G,A).

6.46 Example: When G acts on a locally compact Hausdorff space X, it inducesan action on the C∗-algebra C0(X) by αg(f)(x) = f(g−1x), x ∈ X, g ∈ G,f ∈ C0(X). This provides us with many interesting examples:

(a) Let G act trivially on a point x. Then the associated crossed products CorGand Cof G are just C∗r (G) and C∗f (G), respectively.

(b) Suppose that α is the action of G on itself by translation: αg(s) = gs. ThenC0(G) of,α G = C0(G) or,α G ∼= K(L2(G)).

6.47 Theorem: Let α : G → Aut (A) be the action of a locally compact abelian

group on a unital C∗-algebra A. Then there is an action of G on Aor,α G, calledthe dual action.

Universal C∗-algebras. Interesting examples of C∗-algebras are often describedas universal objects given by generators and relations. This has to be done withsome care, though, because some generators and relations cannot be used to con-struct C∗-algebras. This is because the generators and relations have to be realis-able as bounded operators on a Hilbert space. Thus relations, which will usuallybe algebraic relations among the generators and their adjoints, will require a normcondition.

6.48 Definition: Given a set of generators G and relations R, a representationof (G,R) on a Hilbert space H is a map π : G → B(H) such that π(G) are operatorssatisfying the relations R.

6.49 If A is the free ∗-algebra on the relations R then this induces a representation(H, πG) of A on H. Let (G,R) be a set of generators and relations and let Adenote the free algebra on generators G. Suppose that, for every a ∈ A,

p(a) = sup‖πG(a)‖ | πG is a representation of (G,R)

is finite. Then p(a) is a seminorm on on A6.50 Definition: Given a set of generators and relations (G,R) such that thep(a) < ∞ for every a ∈ A, where A and p are defined as above. The universalC∗-algebra of (G,R), written C∗(G | R), is the enveloping C∗-algebra of (A, p).6.51 The universal C∗-algebra A = C∗(G,R) has the following universal property:If a C∗-algebra B contains a set of elements X in one-to-one correspondence withG which also satisifes the relations R, then there is a surjective ∗-homomorphismA→ C∗(X) where C∗(X) is the C∗(X)-subalgebra of B generated by X.

STRUNG – INVITATION TO C∗-ALGEBRAS 52

6.52 Examples: 1. The first example is a nonexample: There is no universal C∗-algebra generated by a self adjoint element. That is, if G = a and R = a = a∗,then there are no representations of (G,R) where p defines a seminorm. The reason,of course, is that p(a) will certainly never be finite.

2. Let G = a, 1 and R = ‖a‖ ≤ 1, a∗ = a, 1 = 1∗ = 12, 1a = a1 = a. ThenC∗(G | R) ∼= C([−1, 1]).

3. Let G = u and R = u∗u = uu∗ = 1. Then the universal C∗-algebra— theuniversal C∗-algebra generated by a unitary—is then isomorphic to C(T). Note

how this is related to the the group C∗-algebra construction for T ∼= Z.

6.53 An important example is the following. Let n ∈ N and let G = ei,j | 1 ≤i, j ≤ n∪1 and R be the relations eijekl = δjkeil, e

∗ii = e2

ii = eii,∑n

i=1 eii = 1The universal C∗-algebra generated by these generators and relations is of coursejust Mn. Thus, whenever we find elements satisfying these relations (nontrivially)in a given C∗-algebra A, we get a copy of Mn sitting inside of A. We call such aset of elements “matrix units”.

In this section, we’re interested in universal algebras that result in C∗-algebrasthat are, on the one hand, very far removed from the UHF algebras we’ve alreadyencountered—they are never stably finite, for example—yet on the other hand bearsome interesting similarities: they are simple and, like some of our UHF algebras,some of the Cuntz algebras have a certain self-absorbing property.

6.54 Definition: Let A be a unital simple C∗-algebra. A is purely infinite if forevery x ∈ A \ 0 there is a, b ∈ A such that axb = 1.

6.55 Definition: Let n ∈ N \ 0 and let G = s1, . . . , sn. Define relations onG by R =

∑nj=1 sjs

∗j = 1, s∗i si = 1, 1 ≤ i ≤ n. Then the universal C∗-algebra on

G subject to R is well-defined and we call C∗(G | R) the Cuntz algebra of type nand denote it by On.

6.56 We can also define a Cuntz algebra of type ∞, denoted O∞, in the obviousway.

Definition: The Cuntz algebra O∞ is the universal C∗-algebra generated by asequence of isometries (si)i∈N such that

∑nj=1 sjs

∗j ≤ 1 for every n ∈ N.

6.57 For k ∈ N, let W nk denote the set of k-tuples (j1, . . . , jk) were ji ∈ 1, . . . , n

if n <∞ and ji ∈ N if n =∞. Let W n∞ = ∪k∈NW n

k .

For µ = (j1, . . . , jk) ∈ W nk we will denote the element sj1sj2 · · · sjk by sµ. We

may also denote 1 by s0. If µ ∈ W nk then the length of µ, written `(µ), will be k.

If µ = 0 then `(µ) = 0.

Lemma: Let µ, ν ∈ W n∞ and let p = sµs

∗µ and q = sνs

∗ν. Then

(i) If `(µ) = `(ν) then s∗µsν = δµν1, and if s∗µsν 6= 0 then sµ = sν and p = q;

STRUNG – INVITATION TO C∗-ALGEBRAS 53

(ii) if `(µ) < `(ν) and s∗µsν 6= 0 then sν = sµsµ′ with µ′ ∈ W n`(ν)−`(µ) and q < p;

(iii) if `(µ) > `(ν) and s∗µsν 6= 0 then sµ = sνsν′ with ν ′ ∈ W n`(µ)−`(ν) and p < q.

6.58 Lemma: If w 6= 0 is a word in si ∪ s∗i then there are unique elementsµ, ν ∈ W n

∞ such that w = sµs∗ν.

6.59 Let F n0 = C and for k > 0 let F n

k = C∗(sµs∗ν | µ, ν ∈ W n

k ). Let F n = ∪k∈NF nk

Proposition: If n < ∞ then F nk∼= Mnk . Moreover F n

k ⊂ F nk+1 so F n ∼= Un∞,

the UHF algebra of type n∞. If n =∞ the F nk∼= K and F n is an AF algebra.

6.60 Let P be the ∗-algebra generated by s1, . . . , sn ∪ s∗1, . . . , s∗n. If w = sµs∗ν

is a word in si ∪ s∗i then

(i) if `(µ)− `(ν) > 0 then w(s∗1)k = sνsµ(s∗1)k ∈ F n`(µ) and thus w ∈ F n

`(µ)sk1;

(ii) if `(µ)− `(ν) < 0 then (s∗1)kw = (s∗1)ksµs∗ν ∈ F n

`(ν) and thus w ∈ sk1F n`(ν);

(iii) if `(µ) = `(ν) then w ∈ F nr = F n

s . Thus, since any a ∈ P is a linearcombination of words, we can write

a =−1∑

i=−N

si1ai + a0 +n∑i=1

aisi1

where the ai ∈ F n. Set Fi(a) = ai.

6.61 Let A be a C∗-algebra and B ⊂ A a C∗-subalgebra. A conditional expectationis a c.p.c map Φ : A → B satisfying Φ(b) = b for every b ∈ B and Φ(b1ab2) =b1Φ(a)b2 for every b1, b2 ∈ B and a ∈ A. (This last requirement is equivalent tosaying that Φ is a left- and right- B module map.) A conditional expectation Φis faithful when Φ(a∗a) = 0 if and only if a = 0. (A google search for faithfulconditional expectation leads to some amusing results.)

6.62 Proposition: The map F0 : P → F n is a faithful conditional expectation.

6.63 Proposition: Suppose that n < ∞ and x ∈ On is nonzero. Then thereexist a, b ∈ On such that axb = 1.

6.64 Proposition: On is simple and purely infinite.

Exercises

6.1 Characterise all finite-dimensional C∗-algebras.

6.2 Let A = lim−→(An, φn) be the inductive limit of a sequence of C∗-algebras and

let B be a C∗-algebra. For every n ∈ N let ψ(n) : An → B be a ∗-homomorphismsatisfying ψ(n+1) φn = ψ(n) and let ψ : A→ B be the induced ∗-homomorphism.Then

(i) ψ is injective if and only if ker(ψ(n)) ⊂ ker(φ(n)) for every n ∈ N, and

STRUNG – INVITATION TO C∗-ALGEBRAS 54

(ii) ψ is surjective if and only if B = ∪∞j=1ψ(n)(An).

6.3 Let (An, φn) be an inductive sequence of C∗-algebras. Let (nk)k∈N ⊂ N be anincreasing sequence of natural numbers. Show that

lim−→(An, φn) ∼= lim−→(Ank , φnk,nk+1).

6.4 Show that the definition of a UHF algebra p is independent of the choice ofUHF sequence (nk)k∈N for p. Thus any UHF algebra is uniquely identified with asupernatural number p

6.5 Show that a UHF algebra is nuclear (see exercise 5.4).

6.6 Let Uq be a UHF algebra of infinite type p. Show that Up ⊗Up ∼= Up. Supposethat p divides q. Show that Up absorbs Uq in the sense that Uq ⊗ Up ∼= Uq.The universal UHF algebra, denoted Q is the UHF algebra associated to thesupernatural number Πp prime p

∞. Thus the universal UHF algebra absorbs allother UHF algebras (including all matrix algebras!).

6.7 Let p = 2∞. The associated C∗-algebra is often called the CAR algebra (whereCAR stands for canonical anticommutation relations. Let X be a compact Haus-dorff space. Show that the C∗-algebra of functions on X taking values in U2∞ , thatis, C(X,U2∞), absorbs U2∞ .

6.8 A unital C∗-algebra A has real rank zero if the invertible self-adjoint elementsare dense in Asa. There is also a notion of real rank n for n ∈ N. Real rank is ageneralisation of covering dimension to C∗-algebras.

(a) Let X be the Cantor set. Show that X has covering dimension 0. Show that[0, 1] has covering dimension 1. Show that C(X) has real rank zero but C([0, 1])does not. (See exercise 4.5.)

(b) Let a ∈ Mn = Mn×n(B), b ∈ Mn×1(C), c ∈ M1×n(C) and d ∈ C. Let ε > 0.Suppose that d is invertible and there is a′ ∈ Mn that is invertible which satisfies‖a′ − (a− bd−1c)‖ < ε. Show that(

a′ + bd−1c bc d

)−1

=

(In 0−d−1c 1

),

(as elements in Mn+1) and∥∥∥∥( a bc d

)−(a′ + bd−1c b

c d

)∥∥∥∥ < ε.

(c) Let A be a unital C∗-algebra. Suppose b ∈ Inv(A) and there is a ∈ Asa with‖a− b‖ < ε. Show that there is b′ ∈ Asa ∩ Inv(A) with ‖a− b′‖ < ε.

(d) Prove that any UHF algebra has real rank zero.

STRUNG – INVITATION TO C∗-ALGEBRAS 55

6.9 We saw that all UHF algebras are simple. Is the same true for AF algebras?Give a proof or counterexample.

7. A very short introduction to classification for simple nuclearC∗-algebras

In this section we will try to sketch the proof of Elliott’s classification of unitalAF algebras by ordered K-theroy and then given an idea of some more recentresearch directions in the classification and structure of C∗-algebras.

Let A be a unital C∗-algebra and let M∞(A) = ∪n∈NMn(A) where Mn(A) isincluded into Mn+1(A) by copying Mn into the top left corner of Mn+1(A), that is

a 7→(a 00 0

).

If p ∈Mn(A) and q ∈Mm(A) the we define p⊕ q ∈Mn+m(A) to be

p⊕ q =

(p 00 q

).

For any p, q ∈ M∞(A), we extend the Murray–von Neumann equivalence bysetting p ∼ q if there is some m,n ∈ N and some v ∈Mm,n such that v∗v = p andvv∗ = q. In the stably finite case, p ∼ q if and only if there is some k ∈ N suchthat p⊕ 1k ∼ q ⊕ 1k.

Let [p] denote the equivalence class of the projection p ∈M∞(A) and let P(A) beall Murray–von Neumann equivalence classes of projections in M∞(A). Since p⊕qis equivalent to q ⊕ p, P(A) has the structure of an abelian semigroup. K0(A) isdefined to be the Grothendieck group of P(A):

K0(A) = [p]− [q] | p, q ∈M∞(A).

Note that, by the Grothendieck construction, [p] − [q] = [p′] − [q′] in K0(A) ifand only if there is r ∈ P(A) such that [p] + [q′] + [r] = [p′] + [q] + [r].

7.1 Exercise: Suppose that p, q ∈ Mn(A) are orthogonal projections, that is,pq = qp = 0. Then p+ q is a projection in Mn(A). Show that [p] + [q] ∼ [p+ q].

7.2 An ordered abelian group is an abelian group G together with a partial orderthat respects the group structure in the sense that if x ≤ y then x+ z ≤ y + z forevery x, y, z ∈ G and such that

G = x ∈ G | 0 ≤ x − x ∈ G | 0 ≤ x.

As in the case for C∗-algebras, we will denote the positive elements of G by G+.

STRUNG – INVITATION TO C∗-ALGEBRAS 56

7.3 A cone in an abelian group G is a subset H such that H+H ⊂ H, G = H−Hand H ∩ (−H) = 0. If G has a partial order, then G+ is a cone. Conversely, ifH is a cone in G, then setting x ≤ y if and only if y − x ∈ H, defines a partialorder on G.

7.4 Theorem: Let A be a unital stably finite C∗-algebra. Then

K0(A)+ := [p] | p ∈ P(A)

is a cone and hence K0(A) is an ordered abelian group.

Proof: That K0(A)+ + K0(A)+ ⊂ K0(A)+ and K0(A) = K0(A)+ − K0(A)+

is immediate. So we just need to show that K0(A)+ ∩ (−K0(A)+) = 0. Ifp⊕ q ∈Mn(A), then [1n] + [p⊕ q] = [1n] and so [1n− p⊕ q] = 1n. There is m ∈ Nsuch that (1n − p⊕ q)⊕ 1m is Murray–von Neumann equivalent to 1n ⊕ 1m, thatis, there exists v ∈Mn+m(A) with v∗v = (1n − p⊕ q)⊕ 1m and vv∗ = 1n+m. SinceA is stably finite, we must have v∗v = 1n+m. Thus m = 0 and p ⊕ q = 0, whichmeans p = 0 and q = 0. So [p] = [q] = 0, which proves the theorem.

7.5 An order unit for an ordered abelian group G is an element u ∈ G+ such that,for every x ∈ G there exists an n ∈ N such that −nu ≤ x ≤ nu.

Proposition: Let A be a unital stably finite C∗-algebra. Then [1A] is an orderunit for K0(A).

Proof: Exercise.

7.6 Corollary: Let A be a unital AF algebra. Then (K0(A), K0(A)+, [1A]) isan ordered abelian group with distinguished order unit.

7.7 If (G,G+, u) and (H,H+, v) are ordered abelian groups with distinguishedorder units, then a unital positive homomorphism is a homomorphism φ : G→ Hsatisfying φ(G+) ⊂ H+ and φ(u) = v. If φ is a group isomorphism and φ−1 is alsoa unital positive homomorphism, then we call φ a unital order isomorphism.

Let A abd B be C∗-algebras and φ : A→ B a ∗-homomorphism. Define

φ∗ : K0(A)→ K0(B)

by setting φ([p]− [q]) = [φ(p)]− [φ(q)]. (It is easy to see that this is well-defined.)We also have that φ(K0(A)+) ⊂ K0(B)+, and if φ is unital, then φ∗([1A]) =[φ(1A)] = [1B] so φ∗ preservers the order unit. If φ : A → B is an isomorphsim,then φ∗ is a unital order isomorphism.

7.8 Theorem: Let F = Mn1 ⊕ · · · ⊕Mnm be a finite dimensional C∗-algebra and

let e(k)ij , 1 ≤ i, j ≤ nk, 1 ≤ k ≤ K, be matrix units for F . The map

φ : (Zm,Zm+ , (n1, . . . , nm))→ (K0(A), K0(A)+, [1A])

STRUNG – INVITATION TO C∗-ALGEBRAS 57

given by

(r1, . . . , rm)→m∑k=1

rk[e(k)11 ]

is a unital order isomorphism.

Proof: If p ∈ P(F ) then there are pk ∈ P(Mnk), 1 ≤ k ≤ m such thatp =

∑mk=1 pk. Each pk is a projection in a matrix algebra over Mnk where the

equivalence classes of projections are determined by their ranks. Thus pk ∼ rk[e(k)11 ]

for some rk ≥ 0 and so p ∼∑m

k+1 rk[e(k)11 ]. It follows that the map φ is surjective

and that φ is positive. It is also clear that φ((n1, . . . , nm)) =∑m

k=1 nk[e(k)11 ] =∑m

k=1[1k] = [1].

Let πk : F → Mnk be the surjection onto Mnk . Suppose φ((r1, . . . , rm)) = 0.

Then∑m

k=1 rk[e(k)11 ] = 0 so for each l ∈ 1, . . . ,m 0 = (πl)∗(

∑mk=1 rk[e

(k)11 ]) = rl[e

(l)11 ].

Thus the direct sum of rk copies of e(l)11 is equivalent to zero. This is only possible

if rl = 0. So φ is injective.

7.9 Definition: A C∗-algebra A has the cancellation property if whenever p, q ∈P(A) and [p] = [q] in K0(A), then p ∼ q, or, equivalently whenever [p] + [r] =[q] + [r] then [p] = [q].

7.10 Proposition: Every AF algebra has the cancellation property.

7.11 Lemma: Let A be C∗-algebra with the cancellation property. q ∈ A be aprojection and p1, . . . , pn ∈ P(A) such that [q] = [p1⊕· · ·⊕pn]. Then there are arepairwise orthogonal projections p′1, . . . , p

′n ∈ A such that q is Murray–von Neumann

equivalent to∑n

i=1 p′i and each p′i is Murray–von Neumann equivalent to pi.

Proof: It suffices to prove that if r ∈Mn(A), p ∈ A and q ∈ A satisfy [r]+[p] = [q]then there is a p′ ∈ A such that p′p = pp′ = 0 and [r] = [p′]. Let s ∈ Mm(A)satisfy [s] = [q]− [p]− [r]. Since A has the cancellation property, s⊕ r ∼ q − p sothere is v ∈M1,m+n such that v∗v = r ⊕ s and vv∗ = q − p. Let p′ = v(r ⊕ 0m)v∗.Then p′ ∈ A and with w = (r ⊕ 0m)1/2v∗ we have w∗w = p′ and w∗w = r ⊕ 0m,so [p′] = [r]. Since p′ = v(r ⊕ 0m)v∗ ≤ v(1n+m)v = vv∗ = q − p, hence also p′ ≤ qthus p′p = pp′ = 0, as required.

7.12 Theorem: Let A = Mn1 ⊕ · · · ⊕Mnm and B be another finite dimensionalC∗-algebra. Suppose that φ : (K0(A), K0(A)+, [1A]) → (K0(B), K0(B)+, [1B]) is aunital positive homomorphism. Then there exists a homomorphism Φ : A → Bsuch Φ∗ = φ. Moreover, Φ is unique up to conjugation by a unitary.

Proof: Let e(l)ij be a set of matrix units for Mnl ⊂ Mn1 ⊕ · · · ⊕Mnm , 0 ≤ l ≤ m.

Denote by 1l the unit of Mnl . Since φ is positive, φ([1l]) = [pl] for some projection

STRUNG – INVITATION TO C∗-ALGEBRAS 58

pl ∈ P(A). Thus

[p1 ⊕ · · · ⊕ pm] = φ(m∑l=1

[1l]) = φ([1a]) = [1B].

Since B has the cancellation property, there are m mutually orthogonal projectionsq1, . . . , qm ∈ B such that

∑ml=1 ql = [1B] and [φ(1l)] = [pl] = [ql] for every 1 ≤ l ≤

m.

Similarly, we have that φ([e(l)11 ]) = [p

(l)11 for some p

(l)11 ∈ P(A). Thus [ql] =

φ([1l]) = φ(nl[e(l)11 ]) = [p

(l)11 ⊕ · · · ⊕ p

(l)11 ]. Since ql ∈ B, there are mutually or-

thogonal q(l)11 , . . . , q

(l)nl,nl ∈ B with [q

(l)ii ] = [p11(l)] = φ([e

(l)11 ]). Since q

(l)ii ∼ q

(l)11 for

each 1 ≤ i ≤ nl, there are vi ∈ B with v(l)i (v

(l)i )∗ = q

(l)ii and (v

(l)i )∗v

(l)i = q

(l)11 . Put

q(l)ij = v

(l)i (v

(l)j )∗.

One can verify that, for each l, q(l)ij satisfy the matrix relations for Mnl . Since the ql

are pairwise orthogonal, this gives a map from the generators of A to B and hence

a map Φ : A → B. Moreover, we get that Φ∗([e(l)11 ]) = [Φ(e

(l)11 ] = [q

(l)11 ] = φ([e

(l)11 ].

Since [e(l)11 ] 1 ≤ l ≤ m generate K0(A), this implies Φ∗ = φ.

Suppose now that Φ,Ψ : A → B are both unital homomorphisms satisfying

Φ∗ = Ψ∗. Let p(l)ij = Φ(e

(l)ij and q

(l)ij = Ψ(e

(l)ij ), 1 ≤ i, j,≤ nl, 1 ≤ l ≤ m. Then

[p(l)ij ] = Φ∗([e

(l)ij ]) = Ψ∗([e

(l)ij ]) = [q

(l)ij ], so p

(l)ij ∼ q

(l)ij . Thus there are vl ∈ B,

1 ≤ l ≤ m satisfying v∗l vl = p(l)11 and vlv

∗l = q

(l)11 . Set

w =∑

l=1m

∑nli=1 q

(l)il wlp

(l)li .

Then w is a unitary and wp(l)ij w

∗ = q(l)ij for every 1 ≤ i, j ≤ nl and 1 ≤ l ≤ m.

Thus Ψ ∗ (elij) = wΦ(e(l)ij )w∗ for every 1 ≤ i, j ≤ nl and 1 ≤ l ≤ m. Since these

elements generate A, we have Ψ = ad (w) Φ.

7.13 Lemma: Let A, B and C be unital stably finite C∗-algebras, with A finite-dimensional. If φ : K0(A) → K0(C) and ψ : K0(B) → K0(C) are positivehomomorphism which satisfy φ(K0(A)+) ⊂ ψ(K0(A)+). The there is a positivehomomorphism ρ : K0(A)→ K0(B) such that ψ ρ = φ.

Proof: Exercise.

7.14 Lemma: Suppose that A = lim−→(An, φn) and B = lim−→(Bn, ψn) with each φn,

ψn injective and there are ∗-homomorphisms αn : An → Bn and βn : Bn → An+1

STRUNG – INVITATION TO C∗-ALGEBRAS 59

making the following diagram commute:

A1φ1 //

α1

A2φ2 //

α2

A3//

α3

· · · // A

B1ψ1 //

β1>>

B2

β2>>

ψ2 // B3//

β2

>>

· · · // B.

Then there are ∗-isomorphisms α : A→ B and β : B → A making

A1φ1 //

α1

A2φ2 //

α2

A3//

α3

· · · // A

α

B1

ψ1 //

β1>>

B2

β2>>

ψ2 // B3//

β2

>>

· · · // B.

β

OO

commute.

7.15 Lemma: Let A = lim−→(An, φn) be an AF algebra and let F be a finite di-

mensional algebra. Suppose that there are positive homomorphisms α : K0(A1)→K0(F ) and γ : K0(F )→ K0(A) such that γ α = φ

(1)∗ . Then there is n ∈ N and a

positive group homomorphism β : K0(F )→ K0(An) such that

K0(A1)(φ1,n)∗//

α1 %%

K0(An)φ(n)∗ // K0(A)

K0(F )

β

OOγ

99

commutes. Moreover, if the maps φn are unital and α([1A]) = [1F ], then alsoβ([1F ]) = [1An ].

7.16 Theorem: [Elliott] Let A and B be unital approximately finite C∗-algebras.Any ∗-isomorphism Φ : A → B induces an order isomorphism of K0-groups, Φ∗ :(K0(A), K0(A)+, [1A])→ (K0(B), K0(B)+, [1B]).

Conversely, if φ : (K0(A), K0(A)+, [1A]) → (K0(B), K0(B)+, [1B]) is an orderisomorphism, then there is a ∗-isomorphism Φ : A→ B satisfying Φ∗ = φ.

Proof: Let (An, ψn)n∈N and (Bn, ρn)n∈N be inductive limit sequences of finite-dimensional C∗-algebras with limits A and B respectively. We may assume thatthe maps ψn and ρn are unital and injective.

Consider the finite dimensional C∗-algebra A1. Since ψ(1) is a unital homo-

morphism, it induces a unital positive map ψ(1)∗ : K0(A1) → K0(A) and thus,

by composition with φ we have φ ψ(1)∗ : K0(A1) → K0(B). We have that

φ ψ(1)∗ (K0(A)+) ⊂ K0(B)+ so for large enough n1, in fact φ ψ(1)

∗ (K0(A)+) ⊂ρ

(n1)∗ K0(Bn1)+. Thus, by Lemma 7.13 there is

α1 : K0(A1)→ K0(Bn1)

STRUNG – INVITATION TO C∗-ALGEBRAS 60

satisfying ρ(n1)∗ α1 = φ ψ1

∗, hence φ−1 ρ(n1)∗ α1 = ψ1

∗ and we may applyLemma 7.15 to find a m1 ∈ N and a map β1 : K0(Bn1)→ K0(Am1) with ψ(m1)β1 =

φ−1 ρ(n1)∗ . Thus φ ψ(m1) β1 = ρ

(n1)∗ so by applying Lemma 7.15 again, we have

n2 > n1 and a map α2 : Am1 → Bn1 such that φ ψ(m1) = ρ(n1)∗ α2.

Continuing the same way, we find n1, n2, n3, . . . and m1,m2,m3, . . . giving thefollowing commutative diagram

K0(A1) //

α1

K0(Am1) //

α2

K0(Am2) //

α3

· · · // K0(A)

φ

K0(Bn1) //

β188

K0(Bn2)

β288

// K0(Bn3) //

β3

::

· · · // K0(B)

φ−1

OO

By Exercise 7.16, the subsequences (Am1 , ψm1,m2) and (Bn1 , ψn1,n2) have induc-tive limit A and B, respectively. Thus we will relabel Amk by Ak and Bnk by Bk

as well as the connecting maps accordingly.

By Theorem there is a ∗-homomorphism σ1 : A1 → B1 such that (σ)∗ = α andτ : B1 → A2 with τ∗ = β By commutativity of the diagram above, we have that(τ1 σ1)∗ = (ψ1)∗ so there is a unitary u2 ∈ A2 such that ad (u)7.17 What about arbitrary unital C∗-algebras? Can they be classified by K0?The answer is no, as soon as one moves to more complicated C∗-algebras, K0

(even as a unital ordered group) is not enough to distinguish two C∗-algebras.For example, suppose that An = C(T, Fn) where Fn is a finite dimensional C∗-algebra and suppose we have ∗-homomorphisms φn : An → An+1. The inductivelimit A = lim(An, φn) is called an AT algebra (“approximately circle” algebra).In the simple unital case, to distinguish two AT algebras, we need to include theK1-group and tracial state space.

If A is a unital C∗-algebra, then K1(A) = K0(SA) where SA = f : [0, 1] →A | f(0) = f(1) = 0 is called the suspension of A. The K1 group can also bedescribed by equivalence classes of unitaries in Mn(A) in a similar manner to theway K0 is defined for projections. AF algebras always have trivial K1, so this iswhy it does not play a role in the classification invariant for AF algebras.

7.18 How far, then, can we get if we throw K1-groups and tracial states into themix? In fact, quite far! But we’ll need a few definitions first.

First of all, it makes sense to look at simple C∗-algebras. We should be ableto classify these before we can say anything in greater generality. Let’s also stickto the separable case. If things are nonseparable, its unlikely that any invariantwill be in any sense computable. As we’ve often seen so far, it is usually easier todeal with unital C∗-algebras. The final thing we need is some sort of “finiteness”condition. The first guess for such a condition was to restrict to those C∗-algebraswith the completely positive approximation property.

STRUNG – INVITATION TO C∗-ALGEBRAS 61

Definition: A C∗-algebra A has the completely positive approximation propertyif, for every finite subset F ⊂ A and every ε > 0, there is a finite-dimensionalC∗-algebra F and completely positive contract maps ψ : A → F and φ : F → Asuch that

‖φ ψ(a)− a‖ < ε for every a ∈ F .

8. Extra material

Multiplier algebras. 8.1 A left multiplier L of A is a bounded linear operatorL : A→ A which satisfies L(ab) = L(a)b for every a, b ∈ A. Similarly one defines aright multiplier R : A→ A as a bounded linear operator satisfying R(ab) = aR(b)for every a, b ∈ A. To define the multiplier algebra of A, we consider pairs of leftand right multipliers (L,R) with the compatibility condition aL(b) = R(a)b forevery a ∈ A. The pair (L,R) is called a double centraliser. We will denote the setof such pairs by M(A) and show that this is a unital C∗-algebra containing A asan essential ideal. An ideal is called essential if it has nonempty intersection withevery other ideal in A. .

8.2 Proposition: Let (L,R) ∈M(A). Then ‖L‖ = ‖R‖, so we define

‖(L,R)‖ := ‖L‖.

Proof: Note that ‖L(b)‖ = supa,‖a‖≤1 ‖aL(b)‖. From this we have

‖L(b)‖ = supa,‖a‖≤1

‖aL(b)‖

= supa,‖a‖≤1

‖R(a)b‖

≤ supa,‖a‖≤1

‖R(a)‖‖b‖

= ‖R‖‖b‖

Thus ‖L‖ = supb,‖b‖≤1 ‖L(b)‖ ≤ ‖R‖. One shows similarly that ‖R‖ ≤ ‖L‖ fromwhich the result follows.

It is easy to check that we can give M(A) the structure of a vector space byviewing it as a closed subspace as B(A) ⊕ B(A). To show that M(A) is a C∗-algebra we need to define the multiplication and adjoint and then check that thenorm above is indeed a C∗-norm.

8.3 Let L : A→ A be a bounded operator. Define L∗ : A→ A by

L∗(a) = L(a∗)∗.

STRUNG – INVITATION TO C∗-ALGEBRAS 62

The adjoint of (L,R) ∈ M(A) is then just (L,R)∗ = (R∗, L∗). For(L1, R1), (L2, R2) ∈M(A), let

(L1, L2)(R1, R2) = (L1L2, R2, R1).

It is not hard to check that, with these operations and the norm defined aboveM(A) is a unital C∗-algebra with unit (id, id). We also have the following.

8.4 Given any a ∈ A, we can define (La, Ra) ∈M(A) by La(b) = ab and Ra(b) = bafor b ∈ A. Proposition: Let A be a C∗-algebra . Then

A→M(A) : a→ (La, Ra)

identifies A an essential ideal inside M(A), which is proper if and only if A isnonunital.

8.5 There are many important uses for the multiplier algebra, but often it is alittle too big and unwieldy. For example, if X is a locally compact Hausdorffspace which is not compact, then M(C0(X)) ∼= C(βX) where βX is the Stone–Cech compactification of X.

More about representations. 8.6 Let (H, π) be a representation of a C∗-algebraA. A vector ξ ∈ H is called cyclic if the linear span of π(a)ξ ∈ H | a ∈ A, whichwe will denote by π(A)ξ (one might think of this as the orbit of ξ under π(A)) isdense in H. If such a ξ exists, then (H, π) is called a cyclic representation.

Theorem: Let φ : A → C be a positive linear functional. The GNS representa-tion associated to φ is cyclic with cyclic vector ξφ, where ξφ is the unique vectorsatisfying

φ(a) = 〈πφ(a), ξφ〉φfor every a ∈ A.

8.7 Two representations (H1, π1) and (H2, π2) of a C∗-algebra A are unitarily equiv-alent if there is a unitary operator u : H1 → H2 such that π2(a) = uπ1(a)u∗ forevery a ∈ A. As the name suggests, this is an equivalence relation on the repre-sentations of A.

8.8 Proposition: Let (H1, π1) be a cyclic representation of A with cyclic vectorξ and let (H2, π2) be a cyclic representation of A with cyclic vector µ. Then thefollowing are equivalent:

(i) (H1, π1) and (H2, π2) are unitarily equivalent with unitary u : H1 → H2

satisfying uξ = µ,(ii) 〈π1(a)ξ, ξ〉 = 〈π2(a)µ, µ〉 for every a ∈ A.

Proof: That (i) =⇒ (ii) is clear. Conversely, suppose that 〈π1(a)ξ, ξ〉 =〈π2(a)µ, µ〉 for every a ∈ A.

STRUNG – INVITATION TO C∗-ALGEBRAS 63

8.9 A representation (H, π) of A is called nondegenerate if the linear span ofπ(a)ξ | a ∈ A, ξ ∈ H, denoted by π(A)(H), is dense in H, or, equivalently, foreach ξ ∈ H \ 0 there is a ∈ A such that π(a)(ξ) 6= 0. Clearly a cyclic represen-tation is nondegenerate, but a nondegenerate representation need not be cyclic.However, we shall see that it is always the direct sum of cyclic representations.

8.10 Theorem: Let A be a C∗-algebra. Any nondegenerate representation (H, π)is the direct sum of cyclic representations.

Proof: Let ξ ∈ H be nonzero and let Hξ = π(A)ξ. By the Kuratowski–Zornlemma, there is a maximal set S ⊂ H\0 such that Hξ, ξ ∈ S are pairwise orthog-onal. Each Hξ is cyclic and π(A)-invariant. Their direct sum is a representation on⊕ξ∈SHξ

∼= ∪ξ∈SHξ ⊂ H. We will show this is all of H. Let η ∈ (∪ξ∈SHξ)⊥. Then

it is easy to check that Hη will be orthogonal to each of Hξ. We have π(a)(η) ∈ Hη

for every a ∈ A and hence by maximality of S, we must have that π(a)(η) = 0 forevery a ∈ A. Since (H, π) is nondegenerate this means η = 0. Hence ∪ξ∈SHξ = H.

8.11 A representation π : A→ B(H) is irreducible if K ⊂ H closed vector subspacewith π(a)K ⊂ K, then K ∈ 0, H.Theorem:

Tensor products for C∗-algebras. 8.12 Defining a norm on the tensor productis a tricky matter since in general there can be more than one. Here, we constructthe minimal tensor product, or spatial tensor product.

We begin with the tensor product of two Hilbert spaces. Let H and K be Hilbertspaces and form the vector space tensor product H ⊗K.

Proposition: Let H and K be Hilbert spaces. Then there is a unique innerproduct on H ⊗K such that

〈ξ ⊗ η, ξ′ ⊗ η′〉 = 〈ξ, ξ′〉H〈η, η′〉K ,

for every ξ, ξ′ ∈ H and η, η′ ∈ K.

Proof:

8.13 The inner product above makes H ⊗ K into a pre-Hilbert space, and sowe complete it to a Hilbert space, denoted H⊗K. The norm on H⊗K satisfies‖x ⊗ y‖ = ‖x‖‖y‖ for x ∈ H and y ∈ K. We use this to define a tensor productof two C∗-algebras.

Let A be a C∗-algebra with universal representation (H, πA) and let B be a C∗-algebra with universal representation (K, πB). Then there is a a unique injective∗-homomorphism φ : A⊗B → B(H⊗K) such that φ(a⊗b) = πA(a)⊗πB(b). Thus

STRUNG – INVITATION TO C∗-ALGEBRAS 64

we may define a C∗-norm on A⊗B by ‖c‖min = ‖φ(c)‖ for c ∈ A⊗B. Note that‖a⊗ b‖min = ‖a‖‖b‖ for every a ∈ A and b ∈ B.

Definition: The minimal (or spatial) tensor product of A and B is given by

A⊗min B := A⊗B‖·‖min.

Index

K0-group, 43On, 52O∞, 52U(A), 15C∗-algebra

abstract, 15concrete, 15enveloping, 40irreducible, 47nondegenerate, 35simple, 20

C∗-equality, 15∗-algebra, 14ea, 20

Adjoint, 14Approximate unit, 25

quasicentral, 26, 37Approximately finite-dimensional (AF)

algebra, 43UHF algebra, 42

Banach algebra, 4

Canonical projection, 12Cauchy–Schwarz inequality, 31Character, 12Character space, 12Commutant, 34

double commmutant, 34Completely positive map, 29Conditional expectation, 53

faithful, 53Convolution, 46Corner, 26Covariant representation, 49Covering dimension, 29, 44Crossed product

full, 51reduced, 51

Cuntz algebra, 52Cuntz equivalence, 29

subequivalence, 29Cyclic

representation, 58vector, 58

Double centraliser, 57

Dual space, 7

Finite C∗-algebra, 43Fourier–Plancheral transform, 48Functional calculus, 19

Gelfand transform, 13Gelfand–Naimark theorem, 18Group

topological, 45unimodular, 46

Group C∗-algebra, 48full, 48reduced, 48

Group actionon a C∗-algebra, 49

Haar measure, 45Hereditary subalgebra, 26Hilbert–Schmidt operators, 9Homomorphism

of Banach algebras, 12Homomorphsim

of Banach algebras, 14

Ideal, 11essential, 57in a C∗-algebra, 15in a Banach algebra, 11maximal, 11modular, 11, 14

Inductive limit, 40universal property, 41

Inductive sequence, 40Integrated form, 50Invariant subspace, 37Inversion, 46Involution, 14Isometry, 43

Linear functionalweak operator and strong operator

continuous, 36extension, 39positive, 30

Matrix units, 52Measure

Haar measure, 4565

STRUNG – INVITATION TO C∗-ALGEBRAS 66

translation-invariant, 45Multiplier, 57Multiplier algebra, 58Murray–von Neumann equivalence, 28, 43

subequivalence, 43

Neumann series, 5Norm

C∗-norm, 15Banach algebra, 4submultiplicative, 4sup norm, 4

Normal element, 14Nuclear, 39Nuclear dimension, 45

Order zero map, 45

Partial isometry, 28Pontryagin dual, 48Positive map, 29, 30Projection, 14

finite, 43infinite, 43

Purely infinite, 52

Real rank, 44Real rank zero, 54

Representation, 33covariant, 49cyclic, 58faithful, 33Gelfand–Naimark–Segal (GNS), 33irreducible, 59left regular, 48nondegenerate, 47, 58unitary, 47universal, 34

Self-adjoint, 14Seminorm, 40Shift

Bilateral shift operator, 10Simple C∗-algebra, 20Spectral mapping theorem, 20Spectral radius, 8Spectrum, 5

nonunital, 9of a commutative Banach algebra, 12

Square root of a positive element, 21

Stable rank, 45Stably finite, 43State, 30

tracial, 30Strong operator topology, 34Supernatural number, 42

infinite type, 42

Topological group, 45Twisted convolution product, 49

Uniformly hyperfinite (UHF) algebra, 42Unitary, 14Unitary equivalence

of elements, 20Unitisation

Banach algebra, 9minimal, 17

von Neumann algebra, 34enveloping, 34

von Neumann double commutant theorem,35

Weak operator topology, 34Weak-∗ topology, 7

STRUNG – INVITATION TO C∗-ALGEBRAS 67

References

1. G. J. Murphy, C∗-algebras and operator theory, Academic Press, London, 1990.2. O Buachalla, Reamonn, Introduction to quantum symmetries: lecture notes, transcribed by

Fatemeh Khosravi, http://reamonnobuachalla.wordpress.com/notes/, 2016.