06.08.2013 Views

Classifying C∗-algebras

Classifying C∗-algebras

Classifying C∗-algebras

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Classifying</strong> C ∗ -<strong>algebras</strong><br />

Klaus Thomsen<br />

Version pr. 18/3 2005<br />

1


Contents<br />

Chapter 1. Fundamentals 5<br />

1. Basics 5<br />

2. Inductive limits of C ∗ -<strong>algebras</strong> 38<br />

3. K0 and the classification of AF-<strong>algebras</strong> 61<br />

4. Choquet simplices and C ∗ -<strong>algebras</strong> 94<br />

5. Ideals and simple C ∗ -<strong>algebras</strong> 128<br />

Chapter 2. Simple AI-<strong>algebras</strong> 163<br />

Chapter 3. The range of the Elliott invariant for simple unital AI-<strong>algebras</strong> 165<br />

3


1.1. Banach <strong>algebras</strong>.<br />

CHAPTER 1<br />

Fundamentals<br />

1. Basics<br />

Definition 1.1. A Banach algebra A is an algebra over the complex numbers<br />

which is also a Banach space in a norm satisfying<br />

ab ≤ ab, a, b ∈ A. (1.1)<br />

A unit in a Banach algebra A is a non-zero element 1 such that 1a = a1 = a, a ∈<br />

A. When A has a unit we say that A is unital.<br />

Example 1.2. Let X be a set and B(X) the set of bounded functions f : X →<br />

C. Then B(X) is an algebra over C in the obvious way (multiplication is defined<br />

pointwise : fg(x) = f(x)g(x), x ∈ X) and we introduce the norm<br />

f = sup {|f(x)| : x ∈ X}<br />

on B(X). In checking that B(X) is a Banach algebra the only non-trivial point is to<br />

show that B(X) is complete in this norm, i.e. is a Banach space. Let’s do that. So<br />

we consider a Cauchy sequence {fn} in B(X). For each x ∈ X the sequence {fn(x)}<br />

is Cauchy in C and therefore convergent by the completeness of C. Let f : X → C<br />

be the function given by f(x) = limn→∞ fn(x). Let ǫ > 0. Because the sequence<br />

{fn} is Cauchy with respect to the norm · we can find N ∈ N such that<br />

|fn(x) − fm(x)| ≤ fn − fm ≤ min {ǫ, 1} (1.2)<br />

for all n, m ≥ N and all x ∈ X. If we let m tend to ∞ in (1.2) we get<br />

|fn(x) − f(x)| ≤ min {ǫ, 1}, n ≥ N, x ∈ X. (1.3)<br />

From (1.3) it follows that |f(x)| ≤ |fN(x)| + 1 ≤ fN + 1 for all x ∈ X, showing<br />

that f ∈ B(X), and that fn − f ≤ ǫ, n ≥ N, showing that limn→∞ fn = f in<br />

B(X). - Thus B(X) is indeed a Banach space and hence a Banach algebra. Note<br />

that B(X) is unital; the unit is the function on X which takes the constant value 1.<br />

Example 1.3. Let V be a complex Banach space. The set B(V ) of bounded<br />

operators on V is a Banach algebra. The product is simply composition of operators<br />

and the norm is the operator norm<br />

A = sup {A(v) : v ≤ 1}.<br />

Example 1.4. If we take V = C n in Example 1.3 we see that the complex n by n<br />

matrices form a Banach algebra where the product is the usual product of matrices.<br />

This Banach algebra will be denoted by Mn. The unit of Mn is the matrix with all<br />

diagonal entries 1 and all other entries 0.<br />

5


6 1. FUNDAMENTALS<br />

Example 1.5. Let L1 (R) denote the space of complex functions on R which are<br />

absolutely integrable with respect to Lebesgue measure. More precisely L1 (R) is the<br />

space of measurable functions f : R → C for which<br />

<br />

|f(t)| dt < ∞ ;<br />

R<br />

but with two such functions being identified when they agree almost everywhere.<br />

L1 (R) is then a Banach algebra with norm<br />

<br />

|f(t)| dt<br />

and with the convolution product<br />

<br />

f ∗ g(x) = g(t)f(x − t) dt .<br />

R<br />

R<br />

Example 1.6. Let X be a set and B a Banach algebra. Let B(X, B) denote the<br />

vector space of functions f : X → B for which sup x∈X f(x) < ∞. Imitating the<br />

constructions from Example 1.2 one sees that B(X, B) is a Banach algebra.<br />

Closed sub<strong>algebras</strong> of Banach <strong>algebras</strong> are clearly Banach <strong>algebras</strong>. By combining<br />

this observation with the previous examples the reader can use his or her<br />

imagination to construct an unlimited reservoir of examples of Banach <strong>algebras</strong>.<br />

For example, the set of continuous functions f : [0, ∞) → Mn for which the limit<br />

limt→∞ f(t) exists in Mn and is an upper triangular matrix (meaning that all entries<br />

below the diagonal are zero), form a Banach algebra in a natural way.<br />

Definition 1.7. Let a be an element of a unital Banach algebra A. Then a is<br />

invertible if there is an element b ∈ A such that ab = ba = 1. The element b with<br />

this property is called the inverse of a and is denoted by a −1 .<br />

The spectrum σ(a) of a is the set<br />

σ(a) = {λ ∈ C : λ1 − a is not invertible }<br />

and ρ(a) = sup {|λ| : λ ∈ σ(a)} is called the spectral radius of a.<br />

Example 1.8. Consider the Banach algebra Mn. A matrix A ∈ Mn is invertible<br />

in the sense of Definition 1.7 if and only if it is invertible as a linear transformation<br />

C n → C n . This happens if and only if ker A = {0}. Thus, for any λ ∈ C, λ1 − A<br />

is invertible if and only if λ is not an eigenvalue of A. Hence the set σ(A) is the set<br />

of eigenvalues of A.<br />

Example 1.9. Let X be a compact Hausdorff space and consider the Banach<br />

algebra C(X) of continuous complex-valued functions on X. (This is a Banach<br />

subalgebra of the Banach algebra B(X) introduced in Example 1.2.) The constant<br />

function 1 is a unit in C(X) and a function f ∈ C(X) is invertible in C(X) if<br />

and only if f(x) = 0 for all x ∈ X; in which case the inverse f −1 is the function<br />

X ∋ x ↦→ 1 . Thus, for any λ ∈ C, λ1 − f is invertible if and only if f(x) = λ for<br />

f(x)<br />

all x ∈ X; i.e. if and only if λ is not in the image of f. Hence the spectrum σ(f) of<br />

f is the image set f(X).<br />

As the two last examples show, the notion of spectrum of a Banach algebra<br />

element is something which generalizes both the notion of eigenvalue for a matrix<br />

and the image set for a continuous function. Our first objective here is to prove that


1. BASICS 7<br />

the spectrum is always non-empty and say a little about its position in C. That the<br />

spectrum is non-empty is a non-trivial fact and we hope the reader will enjoy the<br />

following proof.<br />

The basic source of invertible elements is the following<br />

Lemma 1.10. Let A be a unital Banach algebra and x ∈ A.<br />

1) If x < 1, it follows that 1 − x is invertible with<br />

(1 − x) −1 =<br />

∞<br />

x n .<br />

n=0<br />

2) If 1 − x < 1, it follows that x is invertible with<br />

x −1 =<br />

∞<br />

(1 − x) n .<br />

Proof. We remark first that x 0 = 1 by definition. a) : The estimate<br />

shows that the partial sums of<br />

<br />

n=0<br />

m<br />

x j ≤<br />

j=n<br />

∞<br />

x n<br />

n=0<br />

m<br />

x j<br />

form a Cauchy sequence in A, so that the infinite sum converges by completeness.<br />

Since (1 − x) ∞ n=0 xn = 1 = ∞ n=0 xn (1 − x), this proves a). b) : Substitute x for<br />

1 − x in a). <br />

Lemma 1.11. Let A be a unital Banach algebra and let x ∈ A be an invertible<br />

element. Then every element y ∈ A with x − y < 1<br />

x−1 is also invertible and<br />

<br />

j=n<br />

x −1 − y −1 ≤ x−1 2 x − y<br />

1 − x −1 x − y .<br />

Proof. Note that 1 − x −1 y = x −1 (x − y) ≤ x −1 x − y < 1 so that<br />

x −1 y is invertible by Lemma 1.10. y must therefore also be invertible with y −1 =<br />

(x −1 y) −1 x −1 = ∞<br />

n=0 (1 − x−1 y) n x −1 , so that x −1 − y −1 ≤ x −1 ∞<br />

n=1<br />

x −1 y n = x −1 1−x−1 y<br />

1−1−x −1 y ≤ x−1 2 x−y<br />

1−x −1 x−y<br />

1 −<br />

. <br />

Note that the qualitative statements of Lemma 1.11 are that the invertible elements<br />

form an open subset of A and that the map x → x −1 is continuous on this<br />

subset. In particular we see that for any sequence {xn} in A we have the equivalence<br />

lim<br />

n→∞ xn = 0 ⇔ lim (1 − xn)<br />

n→∞ −1 = 1. (1.4)<br />

Lemma 1.12. Let {αn} be a sequence of real numbers such that αn+m ≤ αn +αm<br />

for all n, m ∈ N. Then the limit limn→∞ 1<br />

n αn exists and equals infn∈N 1<br />

n αn.


8 1. FUNDAMENTALS<br />

Proof. Fix ǫ > 0 and m ∈ N. Choose N ∈ N so large that<br />

2 m 1<br />

max {|α1|,<br />

N 2 |α2|, . . ., 1<br />

m |αm|} < ǫ.<br />

If n ≥ N we write n = lm + r where l ∈ N and r ∈ {1, 2, . . ., m}. Then<br />

Thus<br />

1<br />

n αn ≤ 1<br />

n αlm + 1<br />

n αr ≤ l<br />

n αm + 1<br />

n αr =<br />

lm 1<br />

n m αm + r 1<br />

n r αr = 1<br />

m αm + r 1<br />

n r αr − r 1<br />

n m αm ≤<br />

1<br />

m αm + 2 m 1<br />

max {|α1|,<br />

N 2 |α2|, . . ., 1<br />

m |αm|} < 1<br />

m αm + ǫ.<br />

lim sup<br />

n→∞<br />

1<br />

n αn ≤ sup<br />

n≥N<br />

1<br />

n αn ≤ 1<br />

m αm + ǫ.<br />

If we let ǫ → 0 and take the infimum over all m ∈ N we see that<br />

lim sup<br />

n→∞<br />

1<br />

n αn ≤ inf<br />

n∈N<br />

1<br />

n αn.<br />

Since obviously infn∈N 1<br />

n αn ≤ lim infn→∞ 1<br />

n αn we conclude that limn→∞ 1<br />

n αn = infn∈N 1<br />

n αn.<br />

<br />

Lemma 1.13. Let A be a unital Banach algebra and a ∈ A. There is an element<br />

n.<br />

λ ∈ σ(a) such that |λ| ≥ limn→∞ a n 1<br />

n = infn∈N a n 1<br />

Proof. We prove first that limn→∞ an 1<br />

n = infn∈N an 1<br />

n. If ak = 0 for<br />

some k ∈ N, an = 0 ∀n ≥ k and the conclusion is trivial. So we assume that<br />

an > 0 ∀n ∈ N and set αn = log an. Note that<br />

By Lemma 1.12 limn→∞ αn<br />

n<br />

lim<br />

n→∞ an 1<br />

n = exp( lim<br />

n→∞<br />

αn+k ≤ αn + αk, n, k ∈ N.<br />

αn = infn∈N . It follows that<br />

n<br />

1<br />

n αn) = exp(inf<br />

n∈N<br />

1<br />

n αn) = inf<br />

n∈N an 1<br />

n.<br />

Set r = limn→∞ an 1<br />

n. We will prove that σ(a) contains an element λ with<br />

|λ| ≥ r. If r = 0 it follows that 0 ∈ σ(a). Indeed, if a is invertible we have that<br />

1 = ana−n so that 1 ≤ ana−n for all n ∈ N and this implies that<br />

1 ≤ r lim a<br />

n→∞ −n 1<br />

n,<br />

which is impossible because r = 0. We can therefore assume that r > 0.<br />

Assume to get a contradiction that λ1 − a is invertible whenever |λ| ≥ r. Fix<br />

1<br />

2πi first n ∈ N and let ω = e n. Set Mk = {1, 2, . . ., n}\{k}, k = 1, 2, . . ., n. We will<br />

use the following identities :<br />

n<br />

and<br />

k=1<br />

n<br />

(1 − ω k z) = 1 − z n , z ∈ C,<br />

<br />

(1 − ω i z) = n, z ∈ C.<br />

k=1 i∈Mk


1. BASICS 9<br />

(If the reader find it difficult to prove these algebraic identities, we offer the following<br />

arguments : The identity n<br />

k=1 (1 − ωk z) = 1 − z n follows from the fact that the<br />

two polynomials have the same roots, namely {ω k : k = 1, 2, · · · , n}. The second<br />

identity follows by observing that the left hand side is a polynomial of degree ≤ n−1,<br />

which takes the same value at n distinct points, namely at z = ω k , k = 1, 2, · · · , n.<br />

It must therefore be the constant polynomial, and the constant value, n, is easily<br />

identified by checking for z = 0.)<br />

By assumption 1 − a is invertible when |λ| ≥ r and the above polynomial iden-<br />

λ<br />

tities show that<br />

n<br />

(1 − ω (1.5)<br />

and<br />

k=1<br />

n<br />

<br />

k=1 i∈Mk<br />

k a an<br />

) = 1 −<br />

λ λn Substituting (1.5) into (1.6) yields that<br />

n 1<br />

(1 −<br />

n<br />

an a<br />

λn)(1 − ωk<br />

λ )−1 = 1<br />

k=1<br />

or, alternatively, that (1 − an<br />

λ n) is invertible with<br />

(1 − an<br />

λ n)−1 = 1<br />

n<br />

(1 − ω ia<br />

) = n1. (1.6)<br />

λ<br />

n<br />

(1 − ω<br />

k=1<br />

k a<br />

λ )−1<br />

(1.7)<br />

when |λ| ≥ r. Set K = sup {(λ1 − a) −1 : |λ| ≥ r}. This is a finite number for the<br />

following reason. When |λ| > a+1 we have that (λ1−a) −1 = λ −1 (1− a<br />

λ )−1 =<br />

λ−1 ∞ a<br />

n=0 ( λ )n ≤ |λ| −1 ∞ a<br />

n=0 ( |λ| )n = 1 ≤ 1. Since M = {λ ∈ C : r ≤ |λ| ≤<br />

|λ|−a<br />

a + 1} is a compact set and λ → (λ1 − a) −1 is continuous by Lemma 1.11, we<br />

see that<br />

K ≤ max {1, sup (λ1 − a)<br />

λ∈M<br />

−1 } < ∞.<br />

For each k ∈ N and |λ| ≥ r we have that<br />

(1 − ω ka<br />

r )−1 k a<br />

− (1 − ω<br />

λ )−1 = ω k a( 1<br />

r<br />

ω k a(λ − r)(r1 − ω k a) −1 (λ1 − ω k a) −1<br />

which gives the estimate<br />

1<br />

− )(1 − ωka<br />

λ r )−1 k a<br />

(1 − ω<br />

λ )−1 =<br />

k a<br />

λ )−1 ≤ |λ − r|aK 2<br />

(1 − ω ka<br />

r )−1 − (1 − ω<br />

when |λ| ≥ r. By substituting this estimate into (1.7) we get that<br />

(1 − an<br />

rn)−1 − (1 − an<br />

λn)−1 ≤ |λ − r|aK 2<br />

when |λ| ≥ r. It is important that this estimate is independent of n.<br />

We prove next that<br />

(1.8)<br />

lim<br />

n→∞ 1 − (1 − an<br />

r n)−1 = 0 . (1.9)


10 1. FUNDAMENTALS<br />

Let ǫ > 0. Choose λ ∈ C such that |λ| > r and aK2 |λ − r| ≤ ǫ.<br />

Since |λ| > r,<br />

2<br />

it follows that limn→∞ an<br />

λn <br />

<br />

r<br />

= 0. To see this, let t ∈ , 1 . Since limn→∞<br />

a<br />

|λ| n<br />

λn <br />

1<br />

n =<br />

|λ| −1 limn→∞ a n 1<br />

n = r<br />

|λ| < t, there is an N so large that an 1<br />

n < t n for all n ≥ N.<br />

Hence limn→∞ an<br />

λ n = 0 because limn→∞ t n = 0.<br />

Now an application of (1.4) gives N ∈ N such that<br />

Combined with (1.8) this shows that<br />

1 − (1 − an<br />

λn)−1 ≤ ǫ<br />

, n ≥ N.<br />

2<br />

1 − (1 − an<br />

r n)−1 ≤ 1 − (1 − an<br />

λ n)−1 + (1 − an<br />

λ n)−1 − (1 − an<br />

r n)−1 ≤ ǫ, n ≥ N,<br />

proving (1.9). Combined with (1.4), (1.9) implies that limn→∞ an<br />

rn = 0. Thus<br />

( an rn ) 1<br />

n = r−1 (an) 1<br />

n < 1 for all sufficiently large n. But this is impossible since<br />

r = infn an 1<br />

n. <br />

Theorem 1.14. Let A be a unital Banach algebra and a ∈ A. Then σ(a) is a<br />

non-empty compact subset of C and<br />

ρ(a) = inf<br />

n∈N an 1<br />

n = lim a<br />

n→∞ n 1<br />

n ≤ a.<br />

Proof. If |λ| n > a n for some n ∈ N, λ n 1 − a n = λ n (1 − an<br />

λ n) is invertible by<br />

Lemma 1.10. Since λ n 1 − a n = (λ1 − a) n−1<br />

k=0 λk a n−k−1 , it follows that λ1 − a is<br />

invertible, i.e. λ /∈ σ(a). - This shows that<br />

ρ(a) ≤ inf<br />

n∈N an 1<br />

n = lim a<br />

n→∞ n 1<br />

n.<br />

But Lemma 1.13 implies that limn→∞ a n 1<br />

n ≤ ρ(a), i.e. we have that<br />

ρ(a) = inf<br />

n∈N an 1<br />

n = lim a<br />

n→∞ n 1<br />

n.<br />

Since a n 1<br />

n ≤ a, this identity implies that ρ(a) ≤ a. In particular, ρ(a) is a<br />

bounded subset of C. The invertible elements of A form an open subset of A by<br />

Lemma 1.11 and hence<br />

C\σ(a) = {λ ∈ C : λ1 − a is invertible}<br />

is open in C; i.e. σ(a) is closed and therefore compact. σ(a) is non-empty by Lemma<br />

1.13. <br />

Corollary 1.15. Let A be a unital Banach algebra. If all non-zero a ∈ A are<br />

invertible we have that<br />

A = C1.<br />

Proof. Let a ∈ A. By Theorem 1.14 σ(a) = ∅, so there is a λ ∈ C such that<br />

λ1 −a is not invertible. By assumption this implies that λ1 −a = 0, i.e. a = λ1.


1. BASICS 11<br />

1.2. Abelian C ∗ -<strong>algebras</strong>. An involution in a Banach algebra A is a map<br />

A ∋ a → a ∗ ∈ A such that<br />

1) (a ∗ ) ∗ = a, a ∈ A,<br />

2) (ab) ∗ = b ∗ a ∗ , a, b ∈ A,<br />

3 (a + λb) ∗ = a ∗ + λb ∗ , a, b ∈ A, λ ∈ C.<br />

Definition 1.16. A Banach algebra A with involution ∗ is a C ∗ -algebra when<br />

for all a ∈ A.<br />

aa ∗ = a 2<br />

(1.10)<br />

Welcome to the world of C ∗ -<strong>algebras</strong>. It is my hope that you will get an enjoyable<br />

stay and maybe even return from time to time when the surrounding world becomes<br />

to hard or boring (or both).<br />

Example 1.17. Let X be a set and B(X) the Banach algebra of bounded complex<br />

functions on X, cf. Example 1.1.2. We define an involution ∗ in B(X) by<br />

f ∗ (x) = f(x), x ∈ X.<br />

It is easy to check that ∗ is an involution on B(X). This turns B(X) into a C ∗ -<br />

algebra. The C ∗ -identity (1.10) is easily established:<br />

ff ∗ = sup {|f(x)f(x)| : x ∈ X} = sup {|f(x)| 2 : x ∈ X}<br />

= sup {|f(x)| : x ∈ X} 2 = f 2<br />

Example 1.18. Let H be a Hilbert space and B(H) the algebra of bounded<br />

operators on H. The operation T → T ∗ which associates to T ∈ B(H) its adjoint<br />

operator is an involution, cf. [GKP, Theorem 3.2.3]. Thus B(H) is a Banach algebra<br />

with involution. It follows from [GKP] that B(H) is a <strong>C∗</strong>-algebra. In particular,<br />

this is true when H is the Hilbert space Cn for any n ∈ N. Now B(Cn ) is the same<br />

as the space of complex n by n matrices Mn, so we see that Mn is a <strong>C∗</strong>- algebra. To<br />

be explicit the <strong>C∗</strong>-norm on Mn is given by<br />

<br />

<br />

<br />

n<br />

A = sup { |<br />

i=1<br />

n<br />

j=1<br />

Aijxj| 2 : (x1, x2, . . ., xn) ∈ C n ,<br />

n<br />

|xi| 2 ≤ 1}<br />

when A = (Aij) ∈ Mn. (I regret that there is no simpler formula - but this is in the<br />

nature of things : There exists no algebraic expression which gives A in terms of<br />

its entries.)<br />

The involution on Mn is reflection in the diagonal followed by conjugation of the<br />

entries :<br />

(A ∗ )ij = Aji.<br />

An alternative proof of the fact that Mn is a C ∗ -algebra will follow below; set<br />

A = C in Proposition 1.53.<br />

We begin now to develop the most elementary general properties of C ∗ -<strong>algebras</strong>.<br />

Note that a general C ∗ -algebra is not assumed to be unital. Certainly, not all<br />

C ∗ -<strong>algebras</strong> have a unit as the following example will show.<br />

i=1


12 1. FUNDAMENTALS<br />

Example 1.19. Let X be a locally compact Hausdorff space and denote by<br />

C0(X) the continuous complex valued functions f on X that ’vanish at infinity’, i.e.<br />

satisfies the following condition : For every ǫ > 0 there is a compact subset K ⊆ X<br />

such that |f(x)| < ǫ for all x ∈ X\K. C0(X) is clearly a subalgebra of B(X) and<br />

is globally invariant under the involution ∗. Furthermore, C0(X) is closed (with<br />

respect to the supremum norm · ). This is seen as follows. Let {fn} be a sequence<br />

from C0(X) and f ∈ B(X) such that limn→∞ fn − f = 0. We first show that f<br />

is continuous : Fix x0 ∈ X and ǫ > 0. Choose N ∈ N such that fN − f < ǫ<br />

3 .<br />

Since fN is continuous at x0 there is an open neighbourhood V of x0 such that<br />

|fN(x)−fN(x0)| < ǫ<br />

3 for all x ∈ V . But then |f(x)−f(x0)| ≤ 2ǫ<br />

3 +|fN(x)−fN(x0)| ≤ ǫ<br />

for all x ∈ V , proving that f is continuous (as should be well-known). To check that<br />

f ∈ C0(X), take K ⊆ X compact such that |fN(x)| < ǫ<br />

for all x ∈ X\K. It follows<br />

3<br />

that |f(x)| ≤ ǫ<br />

3 + |fN(x)| < ǫ for all x ∈ X\K. - Thus C0(X) is closed in B(X) as<br />

asserted and we conclude that C0(X) is a <strong>C∗</strong>-algebra with the operations inherited<br />

from B(X); i.e. C0(X) is <strong>C∗</strong>-subalgebra of B(X).<br />

For every x ∈ X there is a function f ∈ C0(X) such that f(x) = 1. This follows<br />

from an appropriate version of Urysohn’s lemma. So if g is a unit in C0(X) we must<br />

have that g(x) = 1 for all x ∈ X, i.e. g must be the constant function 1 (the unit<br />

of B(X)). But the constant function 1 can only ‘vanish at infinity’ if X is compact.<br />

Indeed, if K is a compact subset of X such that 1 < 1 for x ∈ X\K, then X\K<br />

2<br />

must be empty, and thus X = K is compact. Hence we see that C0(X) can only be<br />

unital when X is compact. Conversely, if X is compact every continuous function<br />

on X ‘vanish at infinity’ (for every ǫ > 0 we can use X itself as the required compact<br />

subset) so C0(X) = C(X) and C(X) contains the constant function 1 as a unit. In<br />

other words: C0(X) is unital if and only if X is compact.<br />

Example 1.20. Let X be a locally compact Hausdorff space and A a C ∗ -algebra<br />

(e.g. A = Mn for some n ∈ N). Denote by C0(X, A) the set of continuous functions<br />

f : X → A which ’vanish at infinity’; i.e. satisfies that for every ǫ > 0 there<br />

is a compact subset K ⊆ X such that f(x) ≤ ǫ ∀x ∈ X\K. A repetition of<br />

the arguments from Example 1.19 shows that C0(X, A) is a C ∗ -algebra with the<br />

involution f ∗ (x) = f(x) ∗ and norm<br />

f = sup {f(x) : x ∈ X}.<br />

When X is compact we denote this algebra by C(X, A). The enthusiastic reader<br />

may prove that C0(X, A) is unital if and only if X is compact and A is unital.<br />

Let us now remedy the fact that a C ∗ -algebra A may not have a unit. Denote<br />

by M(A) the set of maps T : A → A (a priori not even linear) for which there is<br />

another map T ∗ : A → A such that<br />

(T(a)) ∗ b = a ∗ T ∗ (b), a, b ∈ A. (1.11)<br />

Lemma 1.21. Every element T of M(A) is a bounded linear operator on A and<br />

T ∗ ∈ M(A) with (T ∗ ) ∗ = T.<br />

Proof. We first observe that the following implication holds for any element<br />

a ∈ A :<br />

ax = 0 ∀x ∈ A ⇒ a = 0. (1.12)<br />

Indeed, if ax = 0 for all x ∈ A, then in particular a 2 = aa ∗ = 0 which implies<br />

that a = 0.


To show that T is linear, let a, b ∈ A and λ ∈ C. Then<br />

1. BASICS 13<br />

(T(λa + b) − λT(a) − T(b)) ∗ x = (T(λa + b) ∗ − λT(a) ∗ − T(b) ∗ )x =<br />

(λa + b) ∗ T ∗ (x) − λa ∗ T ∗ (x) − b ∗ T ∗ (x) = 0<br />

for all x ∈ A, so by (1.12) we have that T(λa + b) = λT(a) + T(b); i.e. T is linear.<br />

That T is bounded follows from the principle of uniform boundedness in the<br />

following way. For each a ∈ A with a ≤ 1, define Sa : A → A by Sa(b) = T(a) ∗ b.<br />

Each Sa is a bounded linear operator (Sa ≤ T(a)) and for every fixed b ∈ A we<br />

have that Sa(b) = T(a) ∗ b = a ∗ T ∗ (b) ≤ T ∗ (b). Therefore [GKP, Theorem<br />

2.2.9] implies that there is an M < ∞ such that Sa ≤ M for all a. Hence<br />

T(a) ∗ b = Sa(b) ≤ M for all a, b ∈ A with a ≤ 1, b ≤ 1. In particular,<br />

whenever a ≤ 1 and T(a) = 0 we have that<br />

∗ T(a)<br />

T(a) = T(a) ≤ M.<br />

T(a)<br />

It follows that T ≤ M. Finally, by applying the involution to the identity T(a) ∗ b =<br />

a ∗ T ∗ (b) we get immediately that T ∗ ∈ M(A) and that in fact (T ∗ ) ∗ = T. <br />

Proposition 1.22. M(A) is a C ∗ -algebra with the involution T ↦→ T ∗ and the<br />

norm T = sup {T(a) : a ≤ 1}. The product in the algebra is the composition<br />

of operators on A and the identity operator is in M(A) (and is therefore a unit of<br />

M(A)).<br />

Proof. We leave it to the reader to check that M(A) is a subalgebra of the<br />

algebra of bounded linear operators on A. That is, you should check that λT +<br />

S, TS ∈ M(A) when T, S ∈ M(A) and λ ∈ C. It is also trivial that the identity<br />

operator is in M(A).<br />

Let us next check the C ∗ -identity. Since TS ≤ T S because we are dealing<br />

with the operator norm, we see immeditaly that T ∗ T ≤ T ∗ T . Let a, b ∈<br />

A, a ≤ 1, b ≤ 1. Then<br />

T(a) ∗ b = a ∗ T ∗ (b) ≤ T ∗ (b) ≤ T ∗ .<br />

When we set b = T(a)<br />

T(a) in this inequality we get that T(a) ≤ T ∗ (when T(a) =<br />

0). The conclusion is that T ≤ T ∗ , T ∈ M(A). When we substitute T ∗ into this<br />

inequality we get the converse version, T ∗ ≤ T , so we have that T ∗ = T .<br />

Now, for any a ∈ A with a ≤ 1 we have the estimate<br />

T(a) 2 = T(a) ∗ T(a) = a ∗ T ∗ T(a) ≤ T ∗ T ,<br />

so by taking the supremum over all such a we conclude that<br />

T 2 ≤ T ∗ T .<br />

All in all we have the following inequalities<br />

T 2 ≤ T ∗ T ≤ T ∗ T = T 2<br />

from which the C ∗ -identity (1.10) follows.<br />

To show that M(A) is complete with respect to the operator norm, let {Tn} be<br />

a Cauchy sequence in M(A). For each a ∈ A, n, m ∈ N, we have that<br />

Tna − Tma ≤ Tn − Tma .


14 1. FUNDAMENTALS<br />

It follows that {Tna} is a Cauchy sequence in A for all a ∈ A. Since T ∗ n − T ∗ m =<br />

(Tn − Tm) ∗ = Tn − Tm for all n, m, we see that also {T ∗ n } is a Cauchy sequence<br />

in M(A). Consequently, {T ∗ na} is a Cauchy sequence in A for all a ∈ A. We can<br />

therefore define maps T, S : A → A by Ta = limn→∞ Tna, Sa = limn→∞ T ∗ na, a ∈ A.<br />

Note that<br />

(Ta) ∗ b = lim<br />

n→∞ (Tna) ∗ b = lim<br />

n→∞ a ∗ T ∗ n b = a∗ Sb<br />

for all a, b ∈ A, proving that T ∈ M(A) (and that T ∗ = S). To show that<br />

limn→∞ Tn = T in M(A), let ǫ > 0. There is a N ∈ N such that<br />

Tna − Tma ≤ Tn − Tm ≤ ǫ (1.13)<br />

for all n, m ≥ N and all a ∈ A with a ≤ 1. If we let m tend to infinity we get<br />

from (1.13) that<br />

Tna − Ta ≤ ǫ<br />

for all n ≥ N and all a ∈ A with a ≤ 1. Hence Tn − T ≤ ǫ for all n ≥ N. <br />

The C ∗ -algebra M(A) is called the multiplier algebra of A.<br />

Definition 1.23. Let ϕ : A → B be a linear map between the C ∗ -<strong>algebras</strong> A<br />

and B. ϕ is a ∗-homomorphism if it is an algebra homomorphism (i.e. satisfies<br />

that ϕ(ab) = ϕ(a)ϕ(b), ∀a, b ∈ A) and respects the involutions in the sense that<br />

ϕ(a) ∗ = ϕ(a ∗ ), a ∈ A. In addition, if ϕ is a bijection, we call it a ∗-isomorphism.<br />

A ∗-homomorphism ϕ : A → B is called unital when A and B are both unital<br />

and ϕ(1) = 1.<br />

It follows from Exercise 1.91 that a ∗-homomorphism between C ∗ -<strong>algebras</strong> is<br />

always of operator norm ≤ 1, and is isometric if and only if it is injective.<br />

When there is a ∗-isomorphism between two C ∗ -<strong>algebras</strong> A and B we say that<br />

they are isomorphic and write<br />

A ≃ B.<br />

Example 1.24. a) If X and Y are sets and r : Y → X an arbitrary map, then r<br />

defines a ∗-homomorphism ϕ : B(X) → B(Y ) by ϕ(f)(y) = f(r(y)), f ∈ B(X), y ∈<br />

Y . Observe that when X and Y are compact topological spaces and r continuous,<br />

then ϕ maps C(X) into C(Y ). ϕ is a ∗-isomorphism if and only if r is a bijection.<br />

b) Let U ∈ Mn be a unitary matrix (i.e. U ∗ U = 1). Then<br />

ϕ(A) = UAU ∗ , A ∈ Mn,<br />

defines a ∗-homomorphism ϕ : Mn → Mn. ϕ is a ∗-isomorphism and because it<br />

maps to and from the same C ∗ -algebra, it is called an automorphism (of Mn.)<br />

c) If X is compact and A unital we can define unital ∗-homomorphisms ϕ :<br />

A → C(X, A) and ψ : C(X, A) → A by ϕ(a)(x) = a, x ∈ X, and ψ(f) = f(x0),<br />

respectively, where x0 is some fixed point in X. Then ψ ◦ ϕ is the identity map on<br />

A, but ϕ is only a ∗-isomorphism in the rather extreme case where X = {x0}.<br />

For any C ∗ -algebra A there is a ∗-homomorphism ϕA : A → M(A) defined in<br />

the following way : For any a ∈ A define Ta : A → A by<br />

Ta(x) = ax, x ∈ A.<br />

Since (Ta(y)) ∗ x = (ay) ∗ x = y ∗ a ∗ x = y ∗ Ta ∗(x) for all x, y ∈ A we see that Ta ∈ M(A)<br />

and that Ta ∗ = Ta ∗. The map ϕA(a) = Ta is easily seen to be a ∗-homomorphism.


1. BASICS 15<br />

Since Ta(x) ≤ ax, x ∈ A, we see immediately that Ta ≤ a. But<br />

Ta(a ∗ ) = aa ∗ = a 2 , so we can conclude that Ta = a which means that<br />

ϕA is an isometry. Thus ϕA(A) is a C ∗ -subalgebra of M(A) which is ∗- isomorphic,<br />

via ϕA, to A.<br />

We will from now on suppress ϕA in the notation and instead simply consider<br />

A as a C ∗ -subalgebra of M(A) when convenient. In particular, when dealing with<br />

an arbitrary (possibly not unital) C ∗ -algebra A the symbol 1 will denote the unit of<br />

M(A). Note that the result of Exercise 1.65 is that A = M(A) if and only if A is<br />

unital.<br />

While M(A) has the unit which A may lack, it is too large for many purposes.<br />

Therefore we set<br />

 = span{1, A} = C1 + A .<br />

It is obvious that  is a ∗-subalgebra of M(A). We claim that it is in fact a <strong>C∗</strong>- subalgebra, i.e. that it is also closed. This is obvious when A is unital because in<br />

this case  = A and there is nothing to prove. The following lemma handles the<br />

case when A is not unital.<br />

Lemma 1.25. Assume that A is not unital. Then  is a unital <strong>C∗</strong>-subalgebra of<br />

M(A) and there is a unital ∗-homomorphism χ : Â → C such that ker χ = A.<br />

Proof. It is straightforward to check that because 1 /∈ A and A is closed in<br />

M(A), the association λ ↦→ infa∈A λ1 − a defines a norm on C. Thus<br />

inf λ1 − a = K|λ|, λ ∈ C,<br />

a∈A<br />

where K = infa∈A 1 − a > 0. So if {λn} and {an} are sequences in C and A,<br />

respectively, such that λn1+an → x in M(A), then K|λn −λm| = infa∈A λn −λm +<br />

a ≤ λn + an − λm − am for all n, m ∈ N, showing that {λn} is Cauchy and hence<br />

convergent in C, say to λ. But then an = λn1 + an − λn1 converges, say to a ∈ A.<br />

It follows that x = λ1 + a ∈ Â. Thus  is closed in M(A) and is a <strong>C∗</strong>-subalgebra. Since 1 /∈ A, we can define χ :  → C by χ(λ1 + a) = λ, λ ∈ C, a ∈ A. It is<br />

straightforward to check that χ is a ∗-homomorphism. <br />

Note that Lemma 1.25 shows that A is a two-sided ideal in Â, being the kernel<br />

of a homomorphism. (This follows also from Exercise 1.65.)<br />

If now A is an arbitrary <strong>C∗</strong>-algebra and a ∈ A, we set<br />

σ(a) = {λ ∈ C : λ1 − a is not invertible in Â}.<br />

When A is unital  = A, this definition agrees with the one introduced in Section<br />

1.1 (considering A as a Banach algebra only). In particular all the results about σ(a)<br />

obtained in the last section apply.<br />

Definition 1.26. Let A be a C ∗ -algebra and a ∈ A. Then<br />

1) a is normal when a ∗ a = aa ∗ .<br />

2) a is selfadjoint when a ∗ = a.<br />

3) a is unitary when aa ∗ = a ∗ a = 1.<br />

Clearly, a selfadjoint or unitary element is also normal. We will sometimes write<br />

Asa for the set of selfadjoint elements in A.


16 1. FUNDAMENTALS<br />

Lemma 1.27. When a ∈ A is normal,<br />

Proof. We have that<br />

ρ(a) = a.<br />

a 2n<br />

2 = a 2n<br />

(a 2n<br />

) ∗ = (aa ∗ ) 2n<br />

= (aa ∗ ) 2n−1<br />

(aa ∗ ) 2n−1<br />

=<br />

(aa ∗ ) 2n−1<br />

2 = (aa ∗ ) 2n−2<br />

(aa ∗ ) 2n−2<br />

2 = (aa ∗ ) 2n−2<br />

4 = . . .<br />

= aa ∗ 2n<br />

= a 2n+1<br />

,<br />

which implies that a2n = a2n, for all n ∈ N. Thus<br />

ρ(a) = lim a<br />

n→∞ 2n<br />

2−n<br />

= a<br />

in this case, cf. Theorem 1.14. <br />

Lemma 1.28. When a ∈ A is unitary,<br />

σ(a) ⊆ T = {λ ∈ C : |λ| = 1}.<br />

Proof. Note that a is invertible with a −1 = a ∗ so that 0 /∈ σ(a). The formula<br />

λ −1 (λ1 − a)a −1 = −(λ −1 − a −1 ), λ = 0,<br />

shows that λ ∈ σ(a) ⇔ λ −1 ∈ σ(a −1 ). So if λ ∈ σ(a) we have that λ −1 ∈ σ(a ∗ ). Thus<br />

|λ| ≤ a and |λ| −1 = |λ −1 | ≤ a ∗ = a, using Lemma 1.27. Since the unitarity<br />

condition (a ∗ a = aa ∗ = 1) implies that a = a ∗ = 1, this is only possible if<br />

|λ| = 1. <br />

Lemma 1.29. When a ∈ A is selfadjoint,<br />

n=0<br />

σ(a) ⊆ R.<br />

Proof. For each x ∈ A the sum ∞ x<br />

n=0<br />

n<br />

converges in A. We claim that<br />

n!<br />

∞ (x + y) n ∞ x<br />

=<br />

n!<br />

n ∞ y<br />

n!<br />

n<br />

, (1.14)<br />

n!<br />

when xy = yx. To prove this let ǫ > 0. Choose N ∈ N so large that<br />

∞ x<br />

<br />

n ∞ y<br />

n!<br />

n<br />

n! −<br />

m xn m y<br />

n!<br />

n<br />

< ǫ, m ≥ N,<br />

n!<br />

and<br />

n=0<br />

<br />

n=0<br />

∞ (x + y) n<br />

−<br />

n!<br />

n=0<br />

Then m (x+y)<br />

n=0<br />

n<br />

n! = m n=0<br />

<br />

n<br />

j=0<br />

m (x + y) n<br />

−<br />

n!<br />

n=0<br />

m<br />

n=0<br />

n=0<br />

n=0<br />

n=0<br />

n=0<br />

m (x + y) n<br />

< ǫ, m ≥ N.<br />

n!<br />

n=0<br />

n! x<br />

j!(n−j)!<br />

jyn−j n! = m n=0<br />

x n<br />

n!<br />

m<br />

n=0<br />

n x<br />

j=0<br />

j<br />

j!<br />

yn x<br />

= <br />

n!<br />

k,l∈Mm<br />

k y<br />

k!<br />

l<br />

l! ,<br />

y n−j<br />

(n−j)!<br />

and hence<br />

where Mm = {(k, l) ∈ N 2 : m < k + l, k, l ≤ m}. Set L = max{1, x, y}. Since<br />

xk<br />

k!<br />

y l<br />

l!<br />

L2m<br />

≤ [ m when k, l ≤ m and k + l ≥ m, this gives the estimate<br />

]! 2<br />

m (x + y)<br />

<br />

n m x<br />

−<br />

n!<br />

n m<br />

n!<br />

n=0<br />

n=0<br />

n=0<br />

y n<br />

n! ≤ m2<br />

2<br />

L 2m<br />

[ m<br />

2 ]!.


Since limm→∞ m2<br />

2<br />

It follows that<br />

L2m [ m<br />

2<br />

1. BASICS 17<br />

= 0 there is an m ≥ N such that<br />

]!<br />

m (x + y)<br />

<br />

n m x<br />

−<br />

n!<br />

n m y<br />

n!<br />

n<br />

≤ ǫ.<br />

n!<br />

<br />

n=0<br />

∞ (x + y) n<br />

−<br />

n!<br />

n=0<br />

n=0<br />

∞<br />

n=0<br />

x n<br />

n!<br />

n=0<br />

∞<br />

n=0<br />

yn ≤ 3ǫ,<br />

n!<br />

proving (1.14).<br />

Now consider the given selfadjoint element a ∈ A and set eia = ∞ (ia)<br />

n=0<br />

n<br />

. Then n!<br />

(e ia ) ∗ ∞ (−ia)<br />

=<br />

n<br />

.<br />

n!<br />

n=0<br />

Hence (1.14) shows that (eia ) ∗eia = eia (eia ) ∗ = 1 , i.e. eia is a unitary.<br />

Let λ ∈ σ(a). Set zn = n−1 k=0 (iλ)k (ia) n−k−1 . Then zn ≤ nLn where L =<br />

max{1, |λ|, a} so that<br />

n=1<br />

n=1<br />

∞<br />

n=1<br />

converges in A to an element x satisfying that<br />

e iλ − e ia ∞ (iλ)<br />

=<br />

n1 − (ia) n ∞ i(λ1 − a)zn<br />

=<br />

= i(λ1 − a)x = ix(λ1 − a).<br />

n!<br />

n!<br />

Since λ1 − a not is invertible, this shows that e iλ − e ia is not invertible. Since e ia is<br />

unitary, this implies that e iλ ∈ T by Lemma 1.28. Hence λ must be real. <br />

The C ∗ -algebra C0(X) of Example 1.19 is clearly Abelian and our next objective<br />

is to prove that any Abelian C ∗ -algebra is of this form.<br />

Definition 1.30. Let A be a ∗-algebra. A non-zero ∗-homomorphism ω : A → C<br />

is called a character of A.<br />

Theorem 1.31. Let A be an Abelian C ∗ -algebra. Then the set X of characters<br />

of A is a locally compact Hausdorff space in the weak ∗ -topology (inherited from A ∗ )<br />

and A is ∗-isomorphic to C0(X) via the isometric map Φ : A → C0(X) given by<br />

zn<br />

n!<br />

Φ(a)(γ) = γ(a), γ ∈ X.<br />

X is compact if and only if A is unital.<br />

Proof. We assert that<br />

a = sup {|γ(a)| : γ ∈ X}, a ∈ A . (1.15)<br />

Note that this equality says that Φ is isometric. To prove this, let γ ∈ X. First we<br />

show that γ admits an extension γ ′ to  which is a character of  such that γ′ (1) = 1.<br />

This is trivial when A is unital because in this case  = A and γ(1)2 = γ(1 2 ) = γ(1)<br />

so that γ(1) ∈ {0, 1}. Note that γ(1) = 0 is impossible because γ = 0. So assume<br />

that A is not unital. We can then define γ ′ (λ1 + z) = λ + γ(z), λ ∈ C, z ∈ A, and<br />

it is trivial to check that γ ′ is in fact a character of Â.<br />

Now, if γ(a) /∈ σ(a), we would have an element b ∈ Â such that b(γ(a)1 −a) = 1.<br />

But then 1 = γ ′ (1) = γ ′ (b)γ ′ (γ(a)1 − a) = γ ′ (b)(γ ′ (a) − γ ′ (a)) = 0, a contradiction.


18 1. FUNDAMENTALS<br />

Hence γ(a) ∈ σ(a). Since γ ∈ X was arbitrary this shows that sup {|γ(a)| : γ ∈<br />

X} ≤ ρ(a). We conclude from Theorem 1.14 that<br />

a ≥ sup {|γ(a)| : γ ∈ X}.<br />

To prove the reverse inequality, we may assume that a = 0. Take t ∈ σ(a) such<br />

that |t| = ρ(a) = a. This is possible because σ(a) is compact. The set (t1 − a) Â<br />

is a ideal in  which does not contain 1. Let I be an ideal in  which contains<br />

(t1 − a) Â but not 1, and is maximal with respect to these properties. Such an ideal<br />

exists by Zorn’s lemma. Then<br />

I is closed in Â. (1.16)<br />

Â/I = C(1 + I). (1.17)<br />

To prove (1.16) and (1.17), note that I is also an ideal in  which contains<br />

(t1 − a) Â. If 1 ∈ I, there would be an element x ∈ I such that 1 − x < 1. But<br />

then x would be invertible in  by Lemma 1.10 and hence 1 = x−1x ∈ I, which<br />

was not the case. Hence 1 /∈ I. By maximality of I we conclude that I = I. It<br />

follows in particular that Â/I is a Banach algebra in the quotient norm : x + I =<br />

inf{x − z : z ∈ I}. Let χ ∈ Â/I be a non-zero element. If χ was not invertible in<br />

Â/I, χÂ/I would define a proper ideal in Â/I and {x ∈  : x+I ∈ χÂ/I} would be<br />

an ideal properly containing I but not 1, contradicting the maximality of I. Thus<br />

χ is indeed invertible and we conclude from Corollary 1.15 that Â/I = C(1 + I).<br />

It follows from (1.17) that  = C1 + I. Define γ(λ1 + z) = λ, λ ∈ C, z ∈ I.<br />

It is straightforward to check that γ is linear and multiplicative. As above we see<br />

that γ(x) ∈ σ(x) for all x ∈ Â. In particular, γ(x) ∈ R when x = x∗ by Lemma<br />

1.29. An arbitrary element x can be written x = x1 + ix2 where x1 = 1<br />

2 (x + x∗ ) and<br />

x2 = 1<br />

2i (x − x∗ ) are selfadjoint. Then<br />

γ(x) = γ(x1) + iγ(x2) = γ(x1) − iγ(x2) = γ(x1 − ix2) = γ(x ∗ ),<br />

proving that γ is a ∗-homomorphism. Since |γ(a)| = |γ(t1−(t1−a))| = |t| = a = 0<br />

we see first that γ|A = 0, and hence that γ|A ∈ X, and then that<br />

a ≤ sup{|γ(a)| : γ ∈ X}.<br />

Hence (1.15) is established.<br />

We prove now that X is locally compact. By (1.15) γ ≤ 1 for all γ ∈ X,<br />

i.e. X is a subset of the unit ball in A ∗ . Let ω0 ∈ X. To show that X is locally<br />

compact it suffices to find an open neighbourhood V of ω0 whose closure is compact.<br />

Since ω0 = 0 there is an element a ∈ A such that ω0(a) > 1. Set V = {ω ∈ X :<br />

|ω(a)| > 1}. Then V is open in X (in the weak ∗ -topology restricted to X) . Since<br />

V ⊆ K = {ω ∈ X : |ω(a)| ≥ 1}, it suffices to prove that K is compact. K is clearly<br />

closed in the weak ∗ -topology; indeed the function f(ω) = ω(a) is continuous on X<br />

by definition of the weak ∗ -topology, so that K = f −1 ({λ ∈ C : |λ| ≥ 1}) is closed.<br />

But K is a subset of the unit ball of A ∗ which is compact in the weak ∗ -topology.<br />

Hence K is compact.<br />

For each a ∈ A, define Φ(a) : X → C by<br />

Φ(a)(ω) = ω(a), ω ∈ X.<br />

Then Φ(a) is continuous on X by definition of the weak ∗ -topology. Furthermore,<br />

for arbitrary ǫ > 0 we can use the argument above to show that {ω ∈ X : |Φ(a)(ω)| ≥


1. BASICS 19<br />

ǫ} is compact in X. Thus Φ(a) ∈ C0(X). It is then straightforward to check that Φ<br />

defines a ∗-homomorphism Φ : A → C0(X). By (1.15) Φ is an isometry.<br />

The surjectivity of Φ follows from the Stone-Weierstrass theorem in the following<br />

way. 1 Since an arbitrary f ∈ C0(X) admits a decomposition f = g1 + ig2 where<br />

g1, g2 ∈ C0(X) are real-valued, it suffices to show that any real-valued element<br />

f ∈ C0(X) is in Φ(Asa). In fact, by writing f = f ∨ 0 − (−f) ∨ 0, we may even<br />

assume that f ≥ 0. Note that Φ(Asa) is a norm closed algebra of real-valued<br />

functions; norm closed because Φ is an isometry and algebra because A is Abelian.<br />

Let ǫ > 0. It suffices to find c ∈ Asa such that Φ(c) − f ≤ 2ǫ. To this end take a<br />

compact subset K ⊆ X such that f(x) < ǫ, x /∈ K. Then<br />

{h ∈ CR(X) : h = Φ(a)|K for some a ∈ Asa}<br />

is an algebra of continuous real functions on K which separates the points of K<br />

and does not vanish at any point of K. By the Stone-Weierstrass theorem there is<br />

therefore an element a ∈ Asa such that |f(x) − Φ(a)(x)| ≤ ǫ, x ∈ K. Choose a<br />

compact subset K1 ⊆ X, K ⊆ K1, such that |Φ(a)(x)| ≤ ǫ, x /∈ K1. By repeating<br />

the previous argument we find b ∈ Asa such that |Φ(b)(x)−f(x)| ≤ ǫ, x ∈ K1. Now,<br />

by Exercise 1.66 we know that<br />

g = (Φ(a) ∧ Φ(b)) ∨ 0<br />

is in Φ(Asa), i.e. g = Φ(c) for some c ∈ Asa. Since |g(x) − f(x)| ≤ 2ǫ, x ∈ X, we<br />

have found the desired c.<br />

By Example 1.19 X is compact iff A is unital. <br />

Theorem 1.31 shows that there is associated a locally compact Hausdorff space<br />

to any abelian C ∗ -algebra, in such a way that the algebra is ∗-isomorphic to the<br />

C ∗ -algebra of continuous functions on the space that vanish at infinity. One must<br />

then necessarily ask which locally compact Hausdorff spaces can occur this way. The<br />

answer is very simple: They all do !<br />

Lemma 1.32. Let X and Y be topological spaces, Y locally compact and Hausdorff.<br />

Let f : X → Y be a function with the property that h ◦ f is continuous for all<br />

h ∈ C0(Y ). It follows that f is continuous.<br />

Proof. Let V ⊆ Y be an open set. By Urysohn’s lemma there is for each x ∈ V<br />

a function h ∈ C0(Y ) such that h(x) = 1, h(y) ≥ 0 for all y ∈ Y and the support of h<br />

is contained in V . It follows that there is a family hα, α ∈ A, of real-valued functions<br />

in C0(Y ) such that V = <br />

α∈A h−1<br />

α (]0, ∞[). By assumption hα ◦ f is continuous for<br />

all α ∈ A. It follows that<br />

f −1 (V ) = f −1<br />

<br />

α∈A<br />

h −1<br />

α (]0, ∞[)<br />

<br />

= <br />

(hα ◦ f) −1 (]0, ∞[)<br />

is open. <br />

1 The Stone-Weierstrass theorem has different forms in different books. Here I shall apply<br />

a real-valued version; specifically, Theorem 7.32 in ’Principles of Mathematical analysis’ by W.<br />

Rudin. If the reader is familiar with other versions, like [GKP, Analysis now, Cor. 4.3.5], the<br />

following arguments can be changed accordingly, and in most cases greatly simplified.<br />

α∈A


20 1. FUNDAMENTALS<br />

Theorem 1.33. Let X be a locally compact Hausdorff space. Then every character<br />

ω of the C ∗ -algebra C0(X) is of the form<br />

ω(f) = f(xω)<br />

for a unique xω ∈ X. The map ω ↦→ xω is a homeomorphism of the character space<br />

of C0(X) onto X.<br />

Proof. Set I = ker ω = {f ∈ C0(X) : ω(f) = 0}, and note that I is a two-sided<br />

ideal in C0(X). Set<br />

F = {x ∈ X : f(x) = 0 ∀f ∈ I}.<br />

Assume to get a contradiction, that F = ∅. Let f ∈ C0(X), and let ǫ > 0. There<br />

is then a compact subset K ⊆ X such that |f(x)| ≤ ǫ for all x ∈ X\K. Since no<br />

point of K is in F, there is for each x ∈ K an element hx ∈ I such that hx(x) = 1.<br />

Let Ux be an open neighbourhood of x with the property that |hx(y) − 1| ≤ ǫ for all<br />

y ∈ Ux. Let {Uxi : i = 1, 2, . . ., N} be a finite subcover of the cover {Ux : x ∈ K} of<br />

K, and let ϕi, i = 1, 2, . . ., N, be non-negative elements of C0(X) such that<br />

i) 0 ≤ N<br />

i=1 ϕi(x) ≤ 1, x ∈ X,<br />

ii) N i=1 ϕi(y) = 1, y ∈ K,<br />

iii) supp ϕi ⊆ Uxi , i = 1, 2, . . ., N.<br />

Then f ′ = N fϕihxi i=1 ∈ I and<br />

|f ′ (y) − f(y)| = |<br />

for all y ∈ X, and<br />

N<br />

i=1<br />

f(y)ϕi(y) (hxi (y) − 1) | ≤<br />

|f ′ (y) − f(y)| ≤ ǫ + |f ′ (y)| ≤ ǫ +<br />

N<br />

|f(y)|ϕi(y)ǫ = |f(y)|ǫ<br />

i=1<br />

N<br />

|f(y)|ϕi(y)(1 + ǫ) ≤ ǫ + ǫ(1 + ǫ)<br />

i=1<br />

for all y ∈ X\K. It follows that f ′ − f ≤ ǫ + fǫ + ǫ(1 + ǫ). Since ǫ > 0 was<br />

arbitrary, and since I is closed, we conclude that f ∈ I. But f was arbitrary, so we<br />

see that I = C0(X), which is impossible because ω = 0. This contradiction shows<br />

that F = ∅.<br />

Assume next that F contains two different points x1, x2. Define Φ : C0(X) → C 2<br />

by Φ(f) = (f(x1), f(x2)). Then Φ is surjective by Urysohn’s lemma. Since I ⊆ ker Φ,<br />

Φ induces a surjection C0(X)/I → C 2 . However, this is impossible because ω<br />

induces a linear injection C0(X)/I → C, so that C0(X)/I is one-dimensional. This<br />

contradiction shows that F consists of exactly one point, xω. Let h0 ∈ C0(X) be a<br />

function such that ω(h0) = 1. Then h = |h0| 2 is a non-negative function with the<br />

same property. It follows that f − ω(f)h ∈ I for all f ∈ C0(X), and hence<br />

f(xω) = ω(f)h(xω) (1.18)<br />

for all f ∈ C0(X). Choose g ∈ C0(X) such that g (xω) = 1, and note that (1.18)<br />

implies that 1 = |g n (xω) | = |ω(g)| n h(xω) for all n ∈ N. This is only possible if<br />

h(xω) = 1 (and |ω(g)| = 1). It follows that f (xω) = ω(f) for all f ∈ C0(X).<br />

The map ω → xω from the character space of C0(X) to X is surjective: If<br />

x ∈ X, µ(f) = f(x) defines a character µ such that xµ = x. Thus if we for a<br />

second let XC0(X) denote the character space of C0(X), we have a bijection δ :<br />

XC0(X) → X given by δ(ω) = xω. If Ψ : C0(X) → C <br />

XC0(X) is the ∗-isomorphism


1. BASICS 21<br />

of Theorem 1.31 we see that Ψ(f) = f ◦ δ. It follows then from Lemma 1.32 that δ<br />

is a homeomorphism.<br />

<br />

Let A be a C ∗ -algebra. For any set F ⊆ A we let C ∗ (F) denote the least C ∗ -<br />

subalgebra of A containing F. Thus C ∗ (F) is the intersection of all C ∗ -sub<strong>algebras</strong><br />

of A which contains F. Alternatively,<br />

C ∗ (F) = span{x1x2 . . .xn : xi ∈ F ∪ F ∗ , i = 1, 2, . . ., n, n ∈ N},<br />

where F ∗ = {a ∗ : a ∈ F }.<br />

Lemma 1.34. Let a be an invertible element of the C ∗ -algebra A. Then a −1 ∈<br />

C ∗ (1, a).<br />

Proof. We consider first the case where a = a∗ . Let b ∈ Â such that ab = ba =<br />

1. Note that b must be self-ajoint since a is. Then C ∗ (a, b) is an Abelian C ∗ -algebra<br />

with unit, so there is a ∗-isomorphism Φ : C ∗ (a, b) → C(X) where X is a compact<br />

Hausdorff space, cf. Theorem 1.31. Set f = Φ(a), g = Φ(b). Since f is invertible in<br />

C(X) (with 1 = g), we see that the image f(X) ⊆ R of X under f does not contain<br />

f<br />

0. Thus f(X) is a compact subset of R not containing 0. We can therefore find a<br />

uniformly on f(X). It follows<br />

sequence {Pn} of polynomials such that Pn(t) → 1<br />

t<br />

that Pn ◦ f → 1<br />

f = g in C(X). Since Pn ◦ f ∈ <strong>C∗</strong> (1, f) = Φ(<strong>C∗</strong> (1, a)), we conclude<br />

that b ∈ <strong>C∗</strong> (1, a). Consider then the general case. Since a is invertible it follows that<br />

a∗a is also invertible (with (a∗a) −1 = a−1 (a−1 ) ∗ )). Since a∗a is selfadjoint we know<br />

that (a∗a) −1 ∈ <strong>C∗</strong> (1, a∗a) ⊆ <strong>C∗</strong> (1, a). It follows that a−1 = (a∗a) −1a∗ ∈ <strong>C∗</strong> (1, a). <br />

Note that it follows from Lemma 1.34 that the spectrum σ(a) of a (considered<br />

as an element of A) is the same as when we consider a as an element of C ∗ (1, a).<br />

Theorem 1.35. Let a be a normal element of the C ∗ -algebra A. Then there is<br />

a ∗-isomorphism Φa : C(σ(a)) → C ∗ (1, a) such that<br />

Φa(idσ(a)) = a,<br />

where idσ(a) denotes the identity map on σ(a) (i.e. idσ(a)(λ) = λ, λ ∈ σ(a)).<br />

Proof. The crucial observation for the proof is that<br />

σ(a) = {ω(a) : ω ∈ X} (1.19)<br />

where X is the space of characters of C ∗ (1, a). To prove (1.19), assume first that<br />

λ ∈ σ(a); i.e. that λ1 − a is not invertible in Â. Since a is normal <strong>C∗</strong> (1, a) is<br />

Abelian, so there is a ∗-isomorphism Φ : C ∗ (1, a) → C(X) by Theorem 1.31. Since<br />

Φ(λ1 − a) = λ1 − Φ(a) is not invertible in C(X), there is an x ∈ X such that<br />

λ = Φ(a)(x). Then C ∗ (1, a) ∋ b → Φ(b)(x) defines a character ω of C ∗ (1, a) such<br />

that ω(a) = λ. Conversely, if λ = ω(a) for some character ω of C ∗ (1, a), then<br />

λ1 − a is not invertible in Â, for if it was it would be invertible already in <strong>C∗</strong> (1, a)<br />

by Lemma 1.34 and this is clearly not possible : If b = (λ1 − a) −1 ∈ C ∗ (1, a) we<br />

would have that 1 = ω(1) = ω(b(λ1 − a)) = ω(b)ω(λ1 − a) = ω(b)(λ − ω(a)) = 0, a<br />

contradiction. Thus λ ∈ σ(a).<br />

Now define ψ : X → σ(a) by ψ(ω) = ω(a). It is clear that ψ is continuous<br />

and that it is surjective by (1.19). If ω1, ω2 ∈ X are two characters such that<br />

ψ(ω1) = ψ(ω2), then ω1(a ∗ ) = ω1(a) = ω2(a) = ω2(a ∗ ). Since ω1 and ω2 also agree


22 1. FUNDAMENTALS<br />

on 1, we see that they agree on any element of the form a n (a ∗ ) m , n, m ∈ N. These<br />

elements span a dense subset of C ∗ (1, a), so by linearity and continuity ω1 = ω2, i.e.<br />

ψ is also injective.<br />

Let ϕ : σ(a) → X be the inverse of ψ. Define Ψ : C ∗ (1, a) → C(σ(a)) by<br />

Ψ(b)(λ) = Φ(b)(ϕ(λ)), λ ∈ σ(a), where Φ : C ∗ (1, a) → C(X) is the ∗-homomorphism<br />

of Theorem 1.31. It is clear that Ψ is a ∗-isomorphism. When λ = ω(a), ω ∈ X, we<br />

have that ϕ(λ) = ω and hence Ψ(a)(λ) = Φ(a)(ω) = ω(a) = λ. Set Φa = Ψ −1 . <br />

Theorem 1.35 defines a ‘function calculus’ on the normal elements of a <strong>C∗</strong>-algebra in the following way : When f : σ(a) → C is a continuous function on the spectrum<br />

σ(a) of the normal element a ∈ A then f(a) will denote Φa(f) ∈ <strong>C∗</strong> (1, a) ⊆ Â.<br />

If f is the restriction to σ(a) of a polynomial λ0 + λ1x + λ2x2 + · · · + λnxn , then<br />

f(a) = λ01 + λ1a + λ2a2 + · · · + λnan . Note that in general, when A has no unit,<br />

f(a) may not lie in A, only in Â. But we have the following.<br />

Lemma 1.36. Let A be a non-unital C ∗ - algebra and a ∈ A a normal element.<br />

When f ∈ C(σ(a)) and f(0) = 0, it follows that f(a) ∈ C ∗ (a) ⊆ A.<br />

Proof. By the Stone-Weierstrass theorem the functions of the form<br />

z ↦→ <br />

λn,mz n z m<br />

0≤n,m≤N<br />

(i.e. polynomials in z and z) are dense in C(σ(a)). Let Qn be a sequence of functions<br />

of this type such that Qn − f → 0. Since f(0) = 0 we have that Qn(0) → 0. So<br />

if we set Pn = Qn − Qn(0) we get a sequence of function of the same form, but<br />

with no constant term (i.e. with λ0,0 = 0), so that Pn − f → 0. Because there<br />

is no constant term in Pn we see that Pn(a) ∈ C ∗ (a) for all n and hence that<br />

f(a) = limn→∞ Pn(a) ∈ C ∗ (a). <br />

Lemma 1.37. Let a ∈ A be a normal element and f : σ(a) → C a continuous<br />

function. Then<br />

σ(f(a)) = f(σ(a)).<br />

Proof. Since Φa is a ∗-isomorphism, it is clear that σ(f(a)) = σ(f), the<br />

spectrum of f in C(σ(a)). But λ1 − f is invertible in C(σ(a)) if and only if<br />

λ1 − f does not take the value 0 on σ(a), i.e. if and only if λ /∈ f(σ(a)). Hence<br />

σ(f(a)) = σ(f) = f(σ(a)). <br />

Example 1.38. Let G be a group. G could be a topological group (like R for<br />

instance), but here we treat G as if it did not have a topology, i.e. as a discrete<br />

group. Let l2 (G) denote the Hilbert space of square summable functions on G, i.e.<br />

l2 (G) consists of the functions f : G → C such that<br />

<br />

<br />

sup |f(g)| 2 <br />

: F ⊆ G, F finite < ∞ .<br />

g∈F<br />

For each g ∈ G define an operator ug on l 2 (G) by<br />

(ugf)(h) = f(g −1 h) , h ∈ G, f ∈ l 2 (G) .


1. BASICS 23<br />

It is then straightforward to check that uguk = ugk and u∗ g = ug−1 for all g, k ∈ G,<br />

ı.e. u is a unitary represenation of G on l 2 (G). The set<br />

<br />

g∈F<br />

λgug : λg ∈ C, g ∈ F , F finite<br />

is a ∗-subalgebra of B(l 2 (G)) and the closure of this set (with respect to the operator<br />

norm on B(l 2 (G))) is then a C ∗ -algebra which we call the reduced group C ∗ -algebra<br />

of G and denote by <strong>C∗</strong> r (G).<br />

It is a simple exercise to show that <strong>C∗</strong> r (G) is Abelian if and only if G is.<br />

1.3. Positivity and Hilbert C ∗ -modules.<br />

Definition 1.39. An element a of a C ∗ -algebra A is called positive when a = a ∗<br />

(i.e. a is selfadjoint) and σ(a) ⊆ R + = [0, ∞[.<br />

Example 1.40. If A = C0(X) where X is a locally compact Hausdorff space,<br />

then an element f ∈ C0(X) is positive if and only if f(x) ≥ 0, x ∈ X. This is an<br />

immediate consequence of the observation that σ(f) = f(X).<br />

If A = Mn, then an element a ∈ Mn is positive if and only if a = a ∗ and all<br />

eigenvalues of a are non-negative. Note that an element a ∈ Mn is not neccesarily<br />

positive just because σ(a) ⊆ R + ; it may not be selfadjoint. This is the case for<br />

example if a is nilpotent : a2 <br />

0 1<br />

= 0, e.g. a = in M2.<br />

0 0<br />

The set of positive elements in the C ∗ -algebra A will be denoted A+.<br />

Lemma 1.41. Let a = a ∗ ∈ A, a ≤ 1. Then a ∈ A+ if and only if 1 −a ≤ 1.<br />

Proof. By Lemma 1.34 a ∈ A+ ⇔ a ∈ C ∗ (1, a)+. So by Theorem 1.35 we must<br />

show that a real-valued function f ∈ C(σ(a)) with f ≤ 1 is non-negative if and<br />

only if 1 − f ≤ 1. And that is trivial. <br />

Lemma 1.42. Let t ∈ R + , a, b ∈ A+. Then ta + b ∈ A+.<br />

Proof. Set c = ta. Then σ(c) = tσ(a) ⊆ [0, ∞[, e.g. by Lemma 1.37. So<br />

c ∈ A+. To prove that c + b ∈ A+ it suffices to show that (2M) −1 (c + b) ∈ A+<br />

where M > c + b. Since c<br />

M<br />

1− c<br />

M<br />

≤ 1 and 1− b<br />

M<br />

≤ 1, b<br />

M<br />

≤ 1 and hence that 1− c+b<br />

2M<br />

<br />

≤ 1 we know from Lemma 1.41 that<br />

) ≤ 1.<br />

= 1<br />

2<br />

(1− c<br />

M<br />

)+ 1<br />

2<br />

(1− b<br />

M<br />

Thus c+b<br />

2M ∈ A+ by Lemma 1.41 and hence also c + b ∈ A+. <br />

In words Lemma 1.42 says that A+ is a convex cone in A. The positive elements<br />

A+ define a partial ordering of Asa in the natural way:<br />

Lemma 1.43. Let a, b ∈ A. Then<br />

a ≤ b ⇔ b − a ∈ A+.<br />

σ(ab) ∪ {0} = σ(ba) ∪ {0}.<br />

Proof. If λ /∈ σ(ba), (λ1−ab)(1+a(λ1−ba) −1 b) = (1+a(λ1−ba) −1 b)(λ1−ab) =<br />

λ1. Thus if λ = 0, λ /∈ σ(ab). This means that C\ (σ(ba) ∪ {0}) ⊆ C\ (σ(ab) ∪ {0}),<br />

or that σ(ab) ∪ {0} ⊆ σ(ba) ∪ {0}. By symmetry we must have equality. <br />

Proposition 1.44. Let a be an element of the C ∗ - algebra A. Then a is positive<br />

if and only if there is an element b ∈ A such that a = b ∗ b.


24 1. FUNDAMENTALS<br />

Proof. If a is positive, the function f(t) = t 1<br />

2 is continuous on σ(a) and b = f(a)<br />

is selfadjoint with b2 = a. (Note that b ∈ A by Lemma 1.36.) For the converse<br />

assume that a = b∗b. Then a is selfadjoint and we can write a = f(a) − g(a) where<br />

f(t) = t ∨ 0 and g(t) = (−t) ∨ 0. Set c = f(a), d = g(a) and note that c, d are<br />

positive by Lemma 1.37 and cd = 0. We will prove that d = 0. Note first that<br />

(bd) ∗bd = d(c − d)d = −d3 ∈ −A+. Next write bd = s + it with s, t ∈ A selfadjoint,<br />

i.e. s = 1<br />

2 (bd + (bd)∗ ) and t = 1<br />

2i (bd − (bd)∗ ). Then (bd)(bd) ∗ = −(bd) ∗ (bd) + 2(s2 +<br />

t2 ) ∈ A+ + A+ ⊆ A+ where we have used Lemma 1.42 and Lemma 1.37. But<br />

σ((bd) ∗ (bd)) ∪ {0} = σ((bd)(bd) ∗ ) ∪ {0} by Lemma 1.43, so we have (bd) ∗ (bd) ∈ A+,<br />

i.e. we have shown that both (bd) ∗ (bd) and −(bd) ∗ (bd) are positive. But this means<br />

that σ((bd) ∗ (bd)) = {0} which implies that (bd) ∗ (bd) = 0 by Lemma 1.27. Hence<br />

−d3 = 0 and d4 = d4 = 0, i.e. d = 0. Thus a = c ∈ A+.<br />

<br />

Lemma 1.45. Let A be a C ∗ -algebra and a, b ∈ Asa two self-adjoint elements.<br />

Then<br />

a) a ≤ a1 in Â.<br />

b) If a ≤ b, then c ∗ ac ≤ c ∗ bc for all c ∈ A.<br />

c) If 0 ≤ a ≤ b, then a ≤ b.<br />

Proof. a) a1 − a is self-adjoint and it follows e.g. from (1.18) that σ(a1 −<br />

a) = a − σ(a). Since σ(a) ⊆ [−a, a] by Theorem 1.14 and Lemma 1.29, we<br />

conclude that a1 − a ∈ A+, i.e. a ≤ a1.<br />

b) By assumption and Proposition 1.44, b − a = d ∗ d for some d ∈ A. Hence<br />

c ∗ bc − c ∗ ac = c ∗ (b − a)c = (dc) ∗ dc ∈ A+.<br />

c) Set a1 = 1 + a, b1 = 1 + b. Then<br />

a1 = ρ(a1) = max{1 + t : t ∈ σ(a)} = 1 + ρ(a) = 1 + a.<br />

Similarly, b1 = 1 + b, so it suffices to show that that a1 ≤ b1. We may<br />

therefore assume that 1 ≤ a ≤ b. Then the spectra of a and b are contained in<br />

[1, ∞). We can therefore make sense of a −1/2 , and we conclude from a) and b) that<br />

ba −1 ≥ a −1/2 ba −1/2 ≥ a −1/2 aa −1/2 = 1.<br />

a<br />

b<br />

But then σ (ba−1 ) ⊆ [1, ∞), and hence σ<br />

Lemma 1.37. It follows that<br />

a<br />

<br />

a<br />

= ρ ≤ 1,<br />

b b<br />

and hence that a ≤ b.<br />

<br />

= σ<br />

<br />

(ba −1 ) −1<br />

⊆]0, 1] by<br />

Definition 1.46. Let A be a C ∗ -algebra. An inner-product A-module is a<br />

linear space E which is a right A-module (compatible with scalar multiplication:<br />

λ(xa) = (λx)a = x(λa), λ ∈ C, x ∈ E, a ∈ A), together with a map<br />

E × E ∋ (x, y) ↦→ 〈x, y〉 ∈ A<br />

such that<br />

i) 〈x, αy + βz〉 = α 〈x, y〉 + β 〈x, z〉,<br />

ii) 〈x, ya〉 = 〈x, y〉a,


iii) 〈x, y〉 ∗ = 〈y, x〉,<br />

iv) 〈x, x〉 ≥ 0,<br />

v) 〈x, x〉 = 0 ⇒ x = 0.<br />

1. BASICS 25<br />

It follows from iii) and i) that the inner product is conjugate linear in the first<br />

variable. If E satisfies all the conditions for an inner-product A-module, except v),<br />

we call E a semi-inner-product A-module.<br />

Proposition 1.47. When E is a semi-inner-product A-module and x, y ∈ E,<br />

then<br />

〈y, x〉 〈x, y〉 ≤ 〈x, x〉 〈y, y〉<br />

in A.<br />

Proof. Let a ∈ A. Then<br />

0 ≤ 〈xa − y, xa − y〉<br />

= a ∗ 〈x, x〉 a − 〈y, x〉a − a ∗ 〈x, y〉 + 〈y, y〉 (by ii) and iii))<br />

≤ 〈x, x〉 a ∗ a − 〈y, x〉a − a ∗ 〈x, y〉 + 〈y, y〉 (by Lemma 1.45)<br />

(1.20)<br />

Assume first that 〈x, x〉 = 0. Then (1.20) gives us the inequality 〈y, x〉a+a ∗ 〈y, x〉 ∗ ≤<br />

〈y, y〉. With a = 〈y, x〉 ∗ , it follows from Lemma 1.45 that 2 〈y, x〉 〈y, x〉 ∗ ≤ 〈y, y〉 ≤<br />

〈y, y〉 1 in this case. The same conclusion holds when x is replaced by tx for any<br />

t ∈ R, i.e. we conclude that 2t 2 〈y, x〉 〈y, x〉 ∗ ≤ 〈y, y〉 1 for all t ∈ R. It follows<br />

that<br />

〈y, y〉 − 2t 2 σ (〈y, x〉 〈y, x〉 ∗ ) ⊆ [0, ∞)<br />

for all t ∈ R. This is only possible when σ (〈y, x〉 〈y, x〉 ∗ ) ⊆] − ∞, 0], i.e. when<br />

〈y, x〉 〈y, x〉 ∗ ≤ 0. Since 〈y, x〉 〈y, x〉 ∗ ≥ 0, we conclude that 〈y, x〉 = 0 when 〈x, x〉 =<br />

0. Hence the lemma holds when 〈x, x〉 = 0, and we can therefore assume that<br />

〈x, x〉 > 0. Working with x replaced by x<br />

√ 〈x,x〉 , we may then assume that<br />

〈x, x〉 = 1. In this case (1.20) yields the inequality<br />

〈y, x〉a + a ∗ 〈x, y〉 ≤ a ∗ a + 〈y, y〉,<br />

valid for all a ∈ A. This gives the desired inequality when we take a = 〈x, y〉. <br />

Given a semi-inner-product A-module E, we set<br />

x = 〈x, x〉 . (1.21)<br />

Proposition 1.48. In any semi-inner-product module over A, we have the inequalities,<br />

〈x, y〉 ≤ xy,<br />

and<br />

x + y ≤ x + y.<br />

Proof. It follows from Proposition 1.47 and c) of Lemma 1.45 that 〈x, y〉 2 ≤<br />

x 2 y 2 , which is the first inequality. The second is a consequence of the first:<br />

x + y 2 = 〈x + y, x + y〉 = 〈x, x〉 + 〈x, y〉 + 〈y, x〉 + 〈y, y〉 <br />

≤ 〈x, x〉 + 〈x, y〉 + 〈y, x〉 + 〈y, y〉 <br />

≤ x 2 + y 2 + 2xy = (x + y) 2 .


26 1. FUNDAMENTALS<br />

Definition 1.49. Let A be a C ∗ -<strong>algebras</strong>. An inner-product A-module E is<br />

called a Hilbert C ∗ -module over A when E is complete in the norm (1.21).<br />

Example 1.50. a) A is a Hilbert C ∗ -module over A in a natural way: The rightmodule<br />

structure is the given one (from the algebra structure of A), and the inner<br />

product is defined by<br />

〈x, y〉 = x ∗ y.<br />

Note that the norm (1.21) is the given C ∗ -algebra norm.<br />

b) When A = C, a Hilbert C ∗ -module over A is the same as a Hilbert space.<br />

c) The direct sum A n = A ⊕ A ⊕ · · · ⊕A of n copies of A is a Hilbert C ∗ -module<br />

over A. The right-module structure is given by<br />

and the inner product is given by<br />

Note that<br />

(x1, x2, . . .,xn)a = (x1a, x2a, . . .,xna),<br />

〈(x1, x2, . . .,xn), (y1, y2, . . .,yn)〉 = x ∗ 1 y1 + x ∗ 2 y2 + · · · + x ∗ n yn.<br />

max<br />

i xi 2 = max<br />

i x ∗ ixi ≤ x ∗ 1x1 + x ∗ 2x2 + · · · + x ∗ nxn ≤ n max<br />

i xi 2 ,<br />

proving that<br />

max<br />

i xi ≤ (x1, x2, . . ., xn) ≤ √ n max<br />

i xi,<br />

for all (x1, x2, . . .,xn) ∈ A n . These inequalities shows that the completeness of A n<br />

in the norm coming from the inner product follows from the completeness of A.<br />

Given a Hilbert C ∗ -module E over A, let LA(E) denote the set of maps T : E →<br />

E with the property that there is another map T ∗ : E → E such that<br />

for all x, y ∈ E.<br />

〈Tx, y〉 = 〈x, T ∗ y〉<br />

Lemma 1.51. Every element T ∈ LA(E) is a linear bounded operator on E,<br />

T ∗ ∈ LA(E) and (T ∗ ) ∗ = T.<br />

Proof. Rewrite the proof of Lemma 1.21. <br />

Proposition 1.52. LA(E) is a C ∗ -algebra with involution T ↦→ T ∗ and the norm<br />

T = sup{Tx : x ∈ E, x ≤ 1}. The product is the composition of operators.<br />

Proof. Rewrite the proof of Proposition 1.22. <br />

We let Mn(A) denote the set of n by n matrices with entries from A. Thus an<br />

element a = <br />

aij is a matrix<br />

⎛<br />

⎞<br />

a11 a12 a13 . . . a1n<br />

⎜a21<br />

a22 a23 . . . a2n⎟<br />

⎜<br />

⎟<br />

⎜a31<br />

a32 a33 . . . a3n⎟<br />

⎜<br />

⎝<br />

.<br />

⎟<br />

. . . .. . ⎠<br />

an1 an2 an3 . . . ann<br />

such that aij ∈ A for all i, j. Mn(A) is a complex vector space in the obvious way<br />

and an algebra with the product


(ab)ij =<br />

and we define an involution in Mn(A) by<br />

1. BASICS 27<br />

n<br />

k=1<br />

aikbkj<br />

(1.22)<br />

(a ∗ )ij = a ∗ ji . (1.23)<br />

It is straightforward to check that these definitions turn Mn(A) into a ∗-algebra.<br />

Set<br />

<br />

n<br />

a = sup <br />

n<br />

x ∗ ixi <br />

≤ 1 (1.24)<br />

for a = (aij).<br />

i,j,k<br />

x ∗ ka∗ 1<br />

ikaijxj 2 : (x1, x2, . . .,xn) ∈ A n , <br />

Proposition 1.53. Mn(A) is a C ∗ -algebra in the norm (1.24).<br />

Proof. The proof is essentially just an interpretation of the horrible expression<br />

defining the norm which allows one to work with it. For a = (aij) ∈ Mn(A), define<br />

a linear operator Ta : A n → A n in the way dictated by linear algebra :<br />

Ta(x) =<br />

n<br />

i=1<br />

a1ixi,<br />

n<br />

a2ixi, . . .,<br />

i=1<br />

n<br />

i=1<br />

anixi<br />

<br />

i=1<br />

, x = (x1, . . .,xn) ∈ A n .<br />

It is straightforward to check that a ↦→ Ta gives us an injective ∗-homomorphism<br />

from Mn(A) into LA (A n ). It follows therefore from Proposition 1.52 that Mn(A) is<br />

a C ∗ -algebra, except for the completeness and non-degeneracy of the norm. These<br />

properties follow from the inequalities<br />

max<br />

ij<br />

aij ≤ a ≤ n 3<br />

2 max<br />

ij<br />

aij (1.25)<br />

for elements a = (aij) ∈ Mn(A).<br />

To prove the first inequality (from left), let x, y ∈ A, x, y ≤ 1, and set<br />

x ′ = (0, 0, . . ., 0, x, 0, . . ., 0), y ′ = (0, 0, . . ., 0, y ∗ , 0, . . ., 0)<br />

where the non-zero entry occurs on the l’th, respectively, k’th coordinate. Then<br />

yaklx = 〈y ′ , Tax ′ 〉 ≤ Ta = a.<br />

If we set x = a∗ kl akl<br />

akl 2, y = a∗ kl<br />

akl , we get the inequality akl ≤ a. Since k, l ∈<br />

{1, 2, . . ., n} were arbitrary this gives the first inequality. To prove the second, let<br />

x = (x1, . . ., xn) ∈ A n with x ≤ 1. Using c) of Lemma 1.45 we find that xj =<br />

x ∗ j<br />

xj 1<br />

2 ≤ n<br />

i=1 x∗ i<br />

xi 1<br />

2 = x ≤ 1 for all j, and hence n<br />

i,j,k x∗ k a∗ ik<br />

n i,j,k x∗k a∗ 1 <br />

2 3<br />

ikaijxj ≤ n maxij aij<br />

2 1<br />

2 = n 3<br />

aijxj 1<br />

2 ≤<br />

2 maxij aij. It follows that<br />

a ≤ n 3<br />

2 maxij aij.<br />

The completeness of Mn(A) in the C ∗ -norm follows straightforwardly : If {a n }<br />

is a Cauchy sequence in Mn(A) , the first inequality show that {a n ij} is a Cauchy<br />

sequence in A so that aij = limn→∞ a n ij exists for all i, j, and the second inequality<br />

shows that limn→∞ a n = a in Mn(A) when a = (aij).


28 1. FUNDAMENTALS<br />

1.4. Projections, partial isometries and finite dimensional C ∗ - <strong>algebras</strong>.<br />

Definition 1.54. A projection in a C ∗ -algebra is a selfadjoint idempotent, i.e.<br />

an element p such that p = p ∗ = p 2 . A partial isometry is an element v such that<br />

vv ∗ is a projection.<br />

and<br />

Lemma 1.55. If v is a partial isometry. Then<br />

vv ∗ v = v, v ∗ vv ∗ = v ∗ ,<br />

v ∗ v is a projection.<br />

Proof. The calculation<br />

(v − vv ∗ v)(v − vv ∗ v) ∗ = vv ∗ − vv ∗ vv ∗ − vv ∗ vv ∗ + vv ∗ vv ∗ vv ∗<br />

= vv ∗ − 2(vv ∗ ) 2 + (vv ∗ ) 3 = 0<br />

shows that v = vv ∗ v. It follows that (v ∗ v) 2 = v ∗ vv ∗ v = v ∗ v, proving that v ∗ v is a<br />

projection, i.e. also that v ∗ is a partial isometry. It follows that v ∗ vv ∗ = v ∗ . <br />

The projection v ∗ v is called the support projection and vv ∗ the range projection<br />

of v. Thus the range projection of v is the support projection of v ∗ and the support<br />

projection of v is the range projection of v ∗ .<br />

Example 1.56. Let X be a compact Hausdorff space. In C(X) the projections<br />

are the functions f ∈ C(X) with f(X) ⊆ {0, 1} and the partial isometries are the<br />

f’s with f(X) ⊆ T ∪ {0}.<br />

Example 1.57. Let H be a Hilbert space and let B(H) be the C ∗ -algebra of all<br />

bounded operators on H. When p is a projection in B(H) the range space pH is a<br />

closed subspace of H and p is the orthogonal projection H → pH ⊆ H corresponding<br />

to the decomposition H = pH ⊕ (pH) ⊥ . Conversely, when V is a closed subspace<br />

of H the orthogonal projection H → V ⊆ H corresponding to the decomposition<br />

H = V ⊕V ⊥ is a projection in B(H). In this way we have a bijective correspondence<br />

between projections in B(H) and closed subspaces in H.<br />

When v is a partial isometry in B(H) we have decompositions H = v ∗ vH ⊕<br />

(v ∗ vH) ⊥ and H = vv ∗ H ⊕ (vv ∗ H) ⊥ . The restriction of v to v ∗ vH maps this closed<br />

subspace isometrically onto vv ∗ H and v annihilates (v ∗ vH) ⊥ , i.e. v is ’partially an<br />

isometry’. The name partial isometry (in the general case) orginates in this example<br />

which also explains the names ’support projection’ and ’range projection’.<br />

Definition 1.58. A set of matrix units in a C ∗ - algebra A is a subset<br />

where N, n1, n2, . . .,nN ∈ N, such that<br />

and<br />

for all d, b, i, j, k, l.<br />

{e d ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N},<br />

e d ije b kl = δd,bδj,ke d il<br />

e d ∗ d<br />

ij = eji (1.26)<br />

(1.27)


1. BASICS 29<br />

Example 1.59. Let N, n1, n2, . . .,nN ∈ N and consider A = Mn1 ⊕ Mn2 ⊕ · · · ⊕<br />

be the matrix in Mnd all of whose entries are zero, except the j’th<br />

} is a set of matrix units in Mnd ,<br />

MnN . Let fd ij<br />

entry of the i’th row where there is a 1. Then {fd ij<br />

the standard matrix units. To get a set of matrix units in A = Mn1 ⊕ · · · ⊕ MnN we<br />

just put the fd ij ’s into their respective coordinates; i.e. we set<br />

e d ij = (0, 0, . . ., 0, fd ij , 0, . . ., 0)<br />

with f d ij on the d’th coordinate. The resulting set {ed ij : i, j = 1, 2, . . ., nd, d =<br />

1, 2, . . ., N} constitute the standard matrix units in A. This set of matrix units is<br />

the universal such set (when n1, n2, . . .,nN are given), in the sense that every other<br />

set of matrix units with these indices are images of this set under some automorphism<br />

of A.<br />

We shall prove the following theorem which characterizes the C ∗ -<strong>algebras</strong> whose<br />

vector space dimension is finite.<br />

Theorem 1.60. A finite dimensional C ∗ -algebra A is ∗-isomorphic to a finite<br />

direct sum of full matrix <strong>algebras</strong>; i.e. there are numbers N, n1, n2, . . ., nN ∈ N such<br />

that A ≃ Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN .<br />

To present the proof it is appropriate to isolate the following lemmas.<br />

Lemma 1.61. A finite dimensional Abelian C ∗ -algebra A is isomorphic to C n ,<br />

where n = dim A.<br />

Proof. By Theorem 1.31 A ≃ C0(X) for some locally compact Hausdorff space<br />

X. If there were n + 1 distinct elements, say x1, x2, . . .,xn+1, in X we could find,<br />

by Urysohn’s lemma, continuous functions fi, i = 1, 2, . . ., n+1, in C0(X) such that<br />

fi(xj) = δi,j for all i, j. Since the fi’s are linearly independent we would have to<br />

conclude that dimA = dim(C0(X)) ≥ n+1, a contradiction. Hence X contains ≤ n<br />

points. Call them y1, y2, . . .,yk. Define C0(X) ∋ f → (f(y1), f(y2), . . ., f(yk)) ∈ C k .<br />

It is straightforward to prove that this is a ∗-isomorphism. Hence A ≃ C k and thus<br />

k = n. <br />

The C ∗ -algebra C n contains a distinguished set of projections, namely the elements<br />

ei, i = 1, 2, . . ., n, forming the standard vector space basis for C n . If A is a<br />

C ∗ -algebra which is isomorphic to C n , then the image of this set of projections in A<br />

will be called the minimal projections in A.<br />

Lemma 1.62. Let A be a non-zero finite dimensional C ∗ - algebra. Then A is<br />

unital.<br />

Proof. Since every selfadjoint element generates an Abelian C ∗ -subalgebra,<br />

such sub<strong>algebras</strong> always exist in abundance in any C ∗ -algebra. Since A is finitedimensional<br />

the set<br />

{dim B : B is an abelian C ∗ -subalgebra of A} (1.28)<br />

is finite. Let D ⊆ A be an abelian C ∗ -subalgebra of A such that dim D is maximal in<br />

the set (1.28). By Lemma 1.61, D ≃ C d where d = dimD, and hence, in particular,<br />

D is unital and we denote the unit of D by p. Set B = {a ∈ A : pa = ap = 0}.<br />

Then B is in fact {0}. To see this, observe that B is a finite dimensional C ∗ -<br />

algebra in itself, and hence, if nonzero, would contain a non-zero projection q, by


30 1. FUNDAMENTALS<br />

Lemma 1.61. But then Cq + D would be an Abelian C ∗ -subalgebra of A properly<br />

containing D since q /∈ D (qp = 0). This contradicts the maximality of dimD,<br />

and B must be zero as asserted. Now for any a ∈ A it is easy to check that<br />

(a − ap) ∗ (a − ap) = (a ∗ − pa ∗ )(a − ap) ∈ B = {0}, i.e. a = ap. Since p = p ∗ we find<br />

by taking adjoints that pa = a for all a ∈ A. Thus p is a unit in A. <br />

We now give the<br />

Proof. Proof of Theorem 1.60 : If A = {0} we take N = 1 and n1 = 0 and<br />

we are done. So assume that A = {0}. Let Z = {x ∈ A : xa = ax for all a ∈ A},<br />

i.e. Z is the center of A. Since A is unital by Lemma 1.62, Z = {0}. Being a finite<br />

dimensional Abelian C ∗ -algebra Z is therefore isomorphic to C N for some N ∈ N,<br />

cf. Lemma 1.61. Let pi, i = 1, 2, . . ., N, be the minimal projections in Z. Note that<br />

N<br />

i=1 pi = 1. For each i, piA is C ∗ -subalgebra of A and the map<br />

a ↦→ (ap1, ap2, . . .,apN)<br />

defines a ∗-isomorphism A ≃ p1A ⊕ p2A ⊕ · · · ⊕ pNA. So now it will suffice to find<br />

nm ∈ N such that pmA ≃ Mnm for each m.<br />

To simplify notation set B = pmA. As above we can conclude that the center of<br />

B is isomorphic to Cd for some d. But if d ≥ 2 we would have a central projection<br />

q in B which was neither 0 nor pm. q would automatically also be central in A<br />

and hence contradict the minimality of pm. Hence d = 1 and the center of B is<br />

C1 (where we use the notation 1 for the unit (= pm) in B = pmA since this is the<br />

algebra in which we now work.) Let D be a maximal Abelian <strong>C∗</strong>-subalgebra of B,<br />

so that D ≃ Cnm for some nm ∈ N, and let q1, q2, . . .,qnm be the minimal non-zero<br />

projections in D. Since 1 = nm i=1 qi we have that<br />

B =<br />

nm<br />

qiBqj. (1.29)<br />

i,j=1<br />

Consider the C ∗ -subalgebra qkBqk for some k ∈ {1, 2, . . ., nm}. We claim that<br />

qkBqk = Cqk. (1.30)<br />

To prove this it suffices to take an arbitrary non-zero selfadjoint a = a ∗ ∈ qkBqk<br />

and show that a ∈ Cqk. So consider the, necessarily finite dimensional, Abelian<br />

C ∗ -algebra C ∗ (a, qk) ⊆ qkBqk. Then C ∗ (a, qk) ≃ C k for some k ∈ N and we must<br />

show that k = 1. Assume not, i.e. assume that k ≥ 2. In that case there is a<br />

projection p ∈ C ∗ (a, qk) such that p /∈ {0, qk}. Then<br />

span {p, q1, q2, . . .,qnm}<br />

is an Abelian C ∗ -subalgebra of B which contains D and has dimension ≥ nm +<br />

1 because {p, q1, q2, . . .,qnm} is a linearly independent set. This contradicts the<br />

maximality of D and hence k = 1, proving (1.30).<br />

Next we claim that<br />

span BqkB = B . (1.31)<br />

To prove this note that span BqkB is a C ∗ - subalgebra of B and also a two-sided<br />

ideal in B. By Lemma 1.62 there is a unit p in span BqkB. For any a ∈ B, we have<br />

that ap ∈ spanBqkB and hence pap = ap. By taking adjoints we see that ap = pa


1. BASICS 31<br />

for all a ∈ B. So p is a non-zero central projection in B and hence p = 1 because<br />

the center of B is C1. But p = 1 implies that span BqkB = B as asserted.<br />

Let i, j ∈ {1, 2, . . .,nm}. Then<br />

there is a non-zero partial isometry wij ∈ B such that qiBqj = Cwij. (1.32)<br />

When wij is a partial isometry as in (1.32), wijw ∗ ij is a non-zero projection qiBqi,<br />

and hence wijw ∗ ij = qi by (1.30). Similarly, w ∗ ij wij = qj, i.e. wij is a partial isometry<br />

with qj as support projection and qi as range projection.<br />

To prove (1.32), let x be a non-zero element of qiBqj. Then xx ∗ is non-zero<br />

element of qiBqi and, according to (1.30), this means that xx∗ = λqi for some<br />

λ ∈ C. But xx∗ ≥ 0 and non-zero, so λ > 0. Set v = λ−1 2x. Then vv∗ = qi. In<br />

particular, v is a partial isometry and hence so is v ∗ by Lemma 1.55, i.e. v ∗ v is a<br />

projection. This element lies in qjBqj = Cqj, so v ∗ v = qj. Thus wij = v is a partial<br />

isometry in B with support projection qj and range projection qi. It remains to<br />

check that qiBqj = Cwij. Since wij ∈ qiBqj we must show that<br />

qiBqj ⊆ Cwij. (1.33)<br />

Let z ∈ qiBqj\{0}. As above we see that z = µv for some partial isometry µ > 0<br />

and some v ∈ B with vv ∗ = qi and v ∗ v = qj. Note that vw ∗ ij ∈ qiBqi. It follows then<br />

from (1.30) that vw ∗ ij = κqi for some κ ∈ C. By using Lemma 1.55 we find that<br />

v = vv ∗ v = vqj = vw ∗ ij wij = κqiwij = κwijw ∗ ij wij = κwij. Hence z = µv = κµwij ∈<br />

Cwij. This proves (1.33) and hence (1.32).<br />

Let {wij : i, j = 1, 2, · · ·nm} be the partial isometries found in (1.32). We set<br />

fij = w ∗ 1iw1j, i, j = 1, 2, . . ., nm.<br />

We find that fijfkl = 0 when j = k because fij ∈ qiBqj and fkl ∈ qkBql. If j = k<br />

we have<br />

fijfkl = w ∗ 1iw1jw ∗ 1jw1l = w ∗ 1iq1w1l = w ∗ 1iw1l = fil.<br />

Since it is obvious that f ∗ ij = fji we have that {fij : i, j = 1, 2, . . ., nm} is a set of<br />

matrix units in B. It follows from (1.32) that fij = µijwij for some µij ∈ C with<br />

|µij| = 1. Thus<br />

qiBqj = Cfij. (1.34)<br />

Let {eij : i, j = 1, 2, . . ., nm} be the standard system of matrix units in Mnm. We<br />

can then define ϕ : Mnm → B as the linear map with the property that ϕ(eij) = fij<br />

for all i, j. It is easy to check that ϕ is a ∗- homomorphism. If (aij) ∈ Mni and<br />

ϕ((aij)) = <br />

i,j aijfij = 0, then 0 = qk<br />

<br />

i,j aijfijql = aklfkl, so that akl = 0 for<br />

all k, l; i.e. ϕ is injective. To prove that ϕ is also surjective we see from (1.29)<br />

that it suffices to show that qiBqj is contained in the image of ϕ for all i, j. But<br />

qiBqj = Cfij by (1.34) so this is trivial. <br />

Example 1.63. Let G be a finite group and consider the reduced group C ∗ -<br />

algebra, C ∗ r (G), introduced in Example 1.38. Since G is finite both l2 (G) and C ∗ r (G)<br />

must be finite dimensional. By Theorem 1.60<br />

C ∗ r(G) ≃ Mn1 ⊕ Mn2 ⊕ · · · ⊕ Mnm .<br />

But which numbers n1, n2, · · · , nm, occur here ? To answer this, let π : G → B(H)<br />

be an irreducible unitary representation of G on the Hilbert space H. We can then


32 1. FUNDAMENTALS<br />

define a ∗-homomorphism ϕπ : <strong>C∗</strong> r (G) → B(H) by<br />

ϕπ( <br />

λgug) = <br />

λgπ(g) .<br />

g∈G<br />

By Exercise 1.67 below, ϕπ maps C ∗ r(G) onto B(H). Let ˆ G denote the set of unitary<br />

equivalence classes of irreducible unitary representations of G. For each χ ∈ ˆ G we<br />

pick an irreducible unitary representation πχ of G representing χ. The dimension of<br />

the Hilbert space Hπχ on which πχ(G) acts is independent of the particular choice<br />

of πχ and we denote this number by dim χ. Define a ∗-homomorphism<br />

by<br />

g∈G<br />

ϕ : C ∗ r(G) → ⊕ χ∈ ˆ G B(Hπχ) ≃ ⊕ χ∈ ˆ G Mdim χ ,<br />

ϕ(x) = (ϕπχ(x)) χ∈ ˆ G .<br />

We claim that ϕ is a ∗-isomorphism. To see this, note that ker ϕπχ is a two-sided<br />

ideal in C ∗ r (G) and, in particular, a finite dimensional <strong>C∗</strong> -algebra in itself. Let pχ be<br />

the unit in ker ϕπχ, cf. Lemma 1.62. Since ker ϕπχ is a two-sided ideal in C ∗ r (G),<br />

pχ is a central projection in C ∗ r(G). Thus (1 −pπχ)C ∗ r(G) is also a two-sided ideal in<br />

C ∗ r (G) and ϕπχ maps (1 − pπχ)C ∗ r (G) ∗-isomorphically onto B(Hπχ). We claim that<br />

(1 − pχ)(1 − pχ1) = 0 when χ = χ1 . (1.35)<br />

Indeed, if this central projection in C ∗ r (G) was non-zero, the image of it, ϕπχ((1 −<br />

pχ)(1 − pχ1)) ∈ B(πχ) would be a projection invariant under πχ(g) for all g ∈ G and<br />

hence ϕπχ((1 − pχ)(1 − pχ1)) = 1 by irredicibility of πχ. Thus (1 − pχ)(1 − pχ1) =<br />

(1 − pχ) since ϕπχ is injective on (1 − pχ)C ∗ r(G). By symmetry (1 − pχ)(1 − pχ1) =<br />

is a ∗-isomorphism<br />

B(Hπχ) → B(Hπχ ) taking πχ(g) to πχ1(g) for all g ∈ G. By Exercise 1.82 below,<br />

1<br />

(1 − pχ1) and hence (1 − pχ) = (1 − pχ1). Then ϕπχ 1 ◦ ϕ −1<br />

πχ<br />

there is a unitary w : Hπχ → Hπχ 1 such that ϕπχ 1 ◦ ϕ −1<br />

πχ (x) = wxw∗ , x ∈ B(Hπχ).<br />

But this implies that πχ and πχ1 are unitarily equivalent, contradicting that χ = χ1<br />

in ˆ G. This proves (1.35). We can now easily deduce that ϕ is surjective; indeed, if<br />

(xχ) ∈ ⊕ χ∈ ˆ G B(Hπχ), we can find yχ ∈ (1 − pχ)C ∗ r (G) such that ϕπχ(yχ) = xχ for<br />

each χ ∈ ˆ G. Then y = <br />

χ∈ ˆ G yχ ∈ <strong>C∗</strong> r (G) and ϕ(y) = x.<br />

To see that ϕ is also injective, let x ∈ ker ϕ. Consider one of the ∗-homomorphisms<br />

ψi : <strong>C∗</strong> r(G) → Mni = B(Cni ) obtained by combining the abstract ∗-isomorphism<br />

<strong>C∗</strong> r (G) ≃ Mn1 ⊕Mn2 ⊕· · ·⊕Mnm with the projection Mn1 ⊕Mn2 ⊕· · ·⊕Mnm → Mni .<br />

Then G ∋ g ↦→ ψi(ug) is a representation G on Cni which is irreducible since ψi maps<br />

<strong>C∗</strong> r(G) onto B(Cni ). This representation is unitarily equivalent to πχ for some χ ∈ ˆ G.<br />

Since πχ(x) = 0, because x ∈ ker χ, we conclude that ψi(x) = 0. This holds for each<br />

i ∈ {1, 2, · · · , m} and hence x = 0. Consequently ϕ is a ∗-isomorphism as asserted.<br />

In short we have proved that<br />

C ∗ r (G) ≃ ⊕ χ∈ ˆ G Mdim χ .<br />

Therefore the numbers, n1, n2, · · · , nm, are the dimensions of the irreducible representations<br />

of G and m is the number of unitary equivalence classes of irreducible<br />

representations of G, i.e. m = | ˆ G|.


and<br />

1.5. Exercises and additional material.<br />

Exercise 1.64. Let A be a C ∗ -algebra. Show that<br />

1. BASICS 33<br />

a ∗ = a<br />

a ∗ a = a 2<br />

for all a ∈ A. Show that if A is unital, then 1 = 1 ∗ and 1 = 1.<br />

Exercise 1.65. Show that the canonical ∗-homomorphism ϕA : A → M(A) is<br />

surjective if and only if A is unital. Show that ϕA(A) is a two-sided ideal in A.<br />

Exercise 1.66. Let A be a norm closed (in the supremum norm) algebra of realvalued<br />

continuous and bounded functions on a topological space X. Let f, g ∈ A.<br />

Show that f ∧ g = min{f, g} ∈ A and f ∨ g = max{f, g} ∈ A. (Hint : f ∨ g =<br />

1<br />

1<br />

(f +g+|f −g|) and f ∧g = (f +g −|f −g|). Approximate t ↦→ |t| by polynomials,<br />

2 2<br />

uniformly on the image of f − g.)<br />

Exercise 1.67. Let H be a Hilbert space and A ⊆ B(H) a C ∗ -subalgebra of<br />

the bounded operators of H. Assume that A is finite dimensional and that the only<br />

non-zero projection in B(H) which is invariant under A (i.e. satisfies that pap = ap<br />

for all a ∈ A) is 1. Prove that A = B(H). (Hints : Note first that the assumption<br />

implies that 1 is the only non-trivial central projection in A. Since A is finite<br />

dimensional this implies that A ≃ Mm for some m ∈ N. Let {eij : i, j = 1, 2, · · · , m}<br />

be a set of matrix units in A. If f ∈ B(H) is a non-zero projection such that<br />

f ≤ e11, then p = m<br />

i=1 ei1fe1i is a projection in B(H) such that pap = ap for all<br />

a ∈ A. Hence p = 1 by assumption, and thus f = e11. But this means that e11 is a<br />

minimal projection in B(H), and it must therefore be a one-dimensional projection.<br />

Similarly, eii is a one-dimensional projection for each i. Since <br />

i eii = 1 this implies<br />

that dim H = m. But then Mm = B(H) because these vector spaces have the same<br />

dimension, namely m 2 .)<br />

Exercise 1.68. Let A be a <strong>C∗</strong>-algebra. We denote by l2(A) the sequences<br />

(a1, a2, a3, . . .) in A with the property that the sequence<br />

n<br />

{<br />

i=1<br />

a ∗ i ai} ∞ n=1<br />

is convergent in A. Show that l2(A) is a Hilbert <strong>C∗</strong>-module over A with an inner<br />

product given by<br />

∞<br />

< (ai) ∞ i=1 , (bi) ∞ i=1 >=<br />

i=1<br />

a ∗ i bi.<br />

Exercise 1.69. Let A = C[0, 1], and set E = {f ∈ A : f(t) = 0, t ≥ 1}.<br />

Show 2<br />

that E is a Hilbert <strong>C∗</strong>-module over A with an inner product given by<br />

< f, g >= f ∗ g.<br />

Exercise 1.70. Let I be any set, and assume that we have a Hilbert C ∗ -module<br />

Ei over A for all i ∈ I. Is it true, that the direct sum<br />

⊕i∈IEi


34 1. FUNDAMENTALS<br />

is a Hilbert <strong>C∗</strong>-module with an inner product given by<br />

<br />

< (ai), (bi) >= lim < ai, bi >,<br />

F<br />

when we take the limit over the net of finite subsets F ⊆ I, ordered by inclusion ?<br />

I think so, but please check if you can make it work.<br />

Exercise 1.71. Consider the C ∗ -algebra Mn (C), and an element m ∈ Mn(C).<br />

a) Show that the spectrum σ(m) of m is the set of eigenvalues of m.<br />

b) Give an example of an element m ∈ M2(C) whose spectum σ(m) is nonnegative,<br />

i.e. σ(m) ⊆ [0, ∞), but which is not positive.<br />

Exercise 1.72. Let X be a locally compact Hausdorff space and consider the<br />

C ∗ -algebra C0(X). Let f ∈ C0(X). Show that<br />

i∈F<br />

σ(f) = f(X).<br />

Exercise 1.73. Let A be unital C ∗ -algebra, and a ∈ Asa a self-adjoint element.<br />

Then<br />

a = inf{t > 0 : −t1 ≤ a ≤ t1}.<br />

Exercise 1.74. Let p and q be projections in the C ∗ -algebra A. Show that<br />

p ≤ q ⇔ pq = p .<br />

Exercise 1.75. A non-zero projection p in a C ∗ -algebra B is minimal when the<br />

following holds: If q ∈ B is a projection such that q ≤ p, then either p = q or q = 0.<br />

Let A be finite dimensional Abelian C ∗ -algebra of dimension n. Let {ei : i =<br />

1, 2, . . ., n} and {fi : i = 1, 2, . . ., n} be sets of minimal projections in A. Show that<br />

there is a permutation σ of {1, 2, . . ., n} such that<br />

ei = fσ(i), i = 1, 2, . . ., n.<br />

Definition 1.76. A continuous linear functional ω ∈ A ∗ is positive when a ∈<br />

A+ ⇒ ω(a) ≥ 0.<br />

Example 1.77. By the Riesz representation theorem every positive functional<br />

on C(X), where X is some compact Hausdorff space, is given by integration with<br />

respect to a regular Borel measure on X.<br />

On Mn the positive functionals are the ones of the form<br />

Mn ∋ a → Tr(ha)<br />

for some positive element h ∈ Mn. The proof of this statement is an exercise.<br />

Lemma 1.78. (Schwarz inequality) Let ω ∈ A ∗ be positive. Then<br />

|ω(a ∗ b)| 2 ≤ ω(a ∗ a)ω(b ∗ b), a, b ∈ A.<br />

Proof. Set < a, b >= ω(a ∗ b). Then < ·, · > defines a quasi-inner product on A<br />

and the desired inequality follows from the Cauchy-Schwarz inequality for this. For<br />

the convenience of the reader we repeat the proof. First we observe that<br />

ω(a ∗ ) = ω(a), a ∈ A. (1.36)


1. BASICS 35<br />

To see this observe first that when a = a ∗ we can write a = a+ − a− where a+, a− ∈<br />

A+, see the proof of Proposition 1.44. Hence ω(a) = ω(a+) − ω(a−) ∈ R in this<br />

case. In general we write a = b + ic where b, c ∈ Asa. Then ω(a) = ω(b) − iω(c) =<br />

ω(b − ic) = ω(a ∗ ), proving (1.36).<br />

Next let a, b ∈ A, λ ∈ C, |λ| = 1, and set c = λa. Then<br />

0 ≤ ω((tc + b) ∗ (tc + b)) = ω(c ∗ c)t 2 + (ω(b ∗ c) + ω(c ∗ b))t + ω(b ∗ b) =<br />

ω(a ∗ a)t 2 + (ω(b ∗ c) + ω(b ∗ c))t + ω(b ∗ b) =<br />

ω(a ∗ a)t 2 + 2Re ω(b ∗ c)t + ω(b ∗ b), t ∈ R.<br />

This is only possible if (Re ω(b ∗ c)) 2 ≤ ω(a ∗ a)ω(b ∗ b), i.e. if (Re λω(b ∗ a)) 2 ≤ ω(a ∗ a)ω(b ∗ b).<br />

Since this holds also when λ is chosen such that λω(b ∗ a) = |ω(b ∗ a)|, we have the<br />

stated inequality. <br />

Lemma 1.79. Let A be a unital C ∗ - algebra and ω ∈ A ∗ . Then ω is positive if<br />

and only if ω = ω(1).<br />

Proof. Assume first that ω is positive. By Lemma 1.78 we have that |ω(a)| 2 ≤<br />

ω(a ∗ a)ω(1) ≤ ωω(1)a 2 for all a ∈ A. Taking the supremum over all a with<br />

a ≤ 1 yields ω 2 ≤ ωω(1), i.e. ω ≤ ω(1). Since the reverse inequality is<br />

trivial we have that ω(1) = ω. Conversely, assume that ω = ω(1). The first<br />

(and main) problem is to show that ω(a) ∈ R when a = a ∗ . To this end write<br />

ω −1 ω(a) = α + iβ, α, β ∈ R. For all γ ∈ R, ω −1 ω(a + iγ1) = α + i(β + γ).<br />

But σ(a) ⊆ [−a, a] so Lemma 1.37 gives that σ(a + iγ1) ⊆ [−a, a] + iγ.<br />

Since a + iγ1 is normal we know that a + iγ1 = ρ(a + iγ1) ≤ a 2 + γ 2 . Since<br />

|ω −1 ω(a + iγ1)| ≥ |β + γ|, we get that |β + γ| ≤ a 2 + γ 2 or β 2 + 2γβ ≤ a 2 .<br />

Since this holds for all γ ∈ R we must have β = 0, i.e. ω(a) = α ∈ R. Finally, for<br />

any non-zero a ∈ A, 1 − a∗ a<br />

a 2 ≤ 1 by Lemma 1.41 and Proposition 1.44, so<br />

|ω −1 ω(1) − a −2 ω −1 ω(a ∗ a)| ≤ 1.<br />

Since ω −1 ω(1) = 1 and a −2 ω −1 ω(a ∗ a) ∈ R this is only possible if ω(a ∗ a) ≥ 0,<br />

i.e. ω is positive. <br />

Proposition 1.80. Let A be a C ∗ -algebra (= 0) and a ∈ A. There exists a<br />

positive functional ω ∈ A ∗ of norm ω = 1 such that ω(a ∗ a) = a 2 .<br />

Proof. By Theorem 1.35 there is a ∗-isomorphism Φa ∗ a : C(σ(a ∗ a)) → C ∗ (1, a ∗ a)<br />

such that Φa ∗ a(idσ(a ∗ a)) = a ∗ a. Set t = ρ(a ∗ a) ∈ σ(a ∗ a) and define ω0 : C ∗ (1, a ∗ a) →<br />

C by ω0(b) = Φa ∗ a −1 (b)(t). Then ω0 is a positive linear functional of norm 1 on<br />

C ∗ (1, a ∗ a) such that ω0(a ∗ a) = ρ(a ∗ a) = a 2 and ω0(1) = 1. By the Hahn-<br />

Banach theorem there is an extension ω ∈ Â∗ of ω0 with ω = ω0 = 1. Since<br />

ω(1) = ω0(1) = 1, Lemma 1.79 shows that ω is positive on Â. Hence the restriction<br />

to A is also positive and ω(a∗a) = ω0(a∗a) = a2 . This shows in particular that<br />

ω’s restriction to A also has norm 1. <br />

Exercise 1.81. A famous theorem from the childhood of <strong>C∗</strong>-<strong>algebras</strong> says that<br />

every <strong>C∗</strong>-algebra is (isometrically) ∗-isomorphic to a norm-closed ∗-invariant subalgebra<br />

of the bounded operators LC(H) of some Hilbert space H. Since A ⊆ Â, it<br />

suffices to show this when A is unital. We will therefore assume that A is unital in<br />

the following. The proof can then be organized to consist of the following steps.


36 1. FUNDAMENTALS<br />

1) Let ω ∈ A ∗ be a positive functional of norm 1. Set Iω = {a ∈ A : ω(a ∗ a) =<br />

0}. Show that Iω is a linear subspace of A which is also a left A-module.<br />

2) Show that the quotient space A/Iω is an inner product space with inner<br />

product<br />

< a + Iω, b + Iω >= ω(a ∗ b).<br />

3) Let Hω be the Hilbert space obtained by completing the inner product space<br />

from 2). Show that there is a ∗-homomorphism πω : A → LC (Hω) such that<br />

πω(a)(b + Iω) = ab + Iω.<br />

4) Show that πω(a) = 0 ⇒ ω(a ∗ a) = 0.<br />

5) Let H be the Hilbert space direct sum<br />

H = ⊕ωHω,<br />

where we take the sum over all positive norm-one functionals ω on A. Show<br />

that there is an injective and isometric ∗-homomorphism π : A → LC(H).<br />

(Hint: Use Proposition 1.50.)<br />

6) Show that A is ∗-isomorphic to a closed ∗-invariant sub-algebra of LC(H).<br />

Exercise 1.82. Let H1 and H2 be two finite dimensional Hilbert spaces and<br />

ψ : B(H1) → B(H2) a ∗-isomorphism. Show that there is a unitary w : H1 → H2<br />

such that ψ(x) = wxw ∗ for all x ∈ B(H1). (Hint : Let {eij : i, j = 1, 2, · · · , dim H1}<br />

be a set of matrix units for B(H1) which spans B(H1). Show that {ψ(eij)} is a set of<br />

matrix units for B(H2) which spans B(H2). For each i, let λi and µi be unit vectors<br />

in the range of eii and ψ(eii), respectively. Show that we can define an operator<br />

v : H1 → H2 such that vλ1 = µ1 and vλi = 0, i = 2, 3, · · · , dim H2 = dim H2, and<br />

that the unitary<br />

will do the job.)<br />

w = <br />

ψ(ei1)ve1i<br />

i<br />

Exercise 1.83. Let Ai, i = 1, 2, . . ., n be C ∗ - <strong>algebras</strong>. Then the direct sum<br />

⊕ n i=1 Ai is an algebra with coordinate-wise multiplication<br />

(a1, a2, . . .,an)(b1, b2, . . .,bn) = (a1b1, . . .,anbn).<br />

Show that ⊕ n i=1 Ai is a C ∗ -algebra with involution<br />

and norm<br />

(a1, a2, . . .,an) ∗ = (a ∗ 1 , a∗ 2 , . . .,a∗ n )<br />

(a1, a2, . . .,an) = max<br />

i<br />

ai.<br />

Show that ⊕ n i=1Ai is unital if and only if each Ai is unital.<br />

Show that ⊕ n i=1Ai is Abelian if and only if each Ai is Abelian.<br />

Exercise 1.84. Let A1, A2, A3, . . . be a sequence of C ∗ -<strong>algebras</strong>. The infinite<br />

product ∞ i=1 Ai is a ∗-algebra with coordinate-wise operations.<br />

Show that<br />

1) A = {(ai) ∈ ∞ i=1 Ai : supi ai < ∞}<br />

2) C = {(ai) ∈ ∞ i=1 Ai : limn→∞ ai = 0 }<br />

are <strong>C∗</strong>-<strong>algebras</strong> in the norm<br />

(ai) = sup<br />

i<br />

ai.


1. BASICS 37<br />

Exercise 1.85. Let A be a ∗-algebra and · 1 and · 2 two norms on A such<br />

that A is a C ∗ -algebra with respect to both norms.<br />

Show that a1 = a2 for all a ∈ A.<br />

In contrast, A has uncountably many Banach algebra norms, e.g. · = k · 1<br />

for any k ≥ 1.<br />

Exercise 1.86. Let A be a C ∗ -algebra. Construct a ∗-homomorphism ϕ : A →<br />

A ⊕ A which is unital when A is. Construct a ∗-homomorphism ψ : A → M2(A)<br />

which is unital when A is.<br />

Let k, n ∈ N. Show that Mk(Mn(A)) is ∗-isomorphic to Mkn(A).<br />

Exercise 1.87. Let X be a locally compact Hausdorff space. Show that<br />

ϕ(A)(x) = (Aij(x)), x ∈ X, A ∈ Mn(C0(X)),<br />

defines a ∗-isomorphism ϕ : Mn(C0(X)) → C0(X, Mn).<br />

Exercise 1.88. Consider the two C ∗ -<strong>algebras</strong> A = Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN and<br />

B = Mm1 ⊕ Mm2 ⊕ · · · ⊕ MmM , where ni ≥ 1 and mj ≥ 1 for all i, j. Show that<br />

A ≃ B if and only if N = M and there is a permuation σ of {1, 2, . . ., N} such that<br />

nσ(i) = mi, i = 1, 2, . . ., N.<br />

Exercise 1.89. Let X and Y be compact Hausdorff spaces and ϕ : C(X) →<br />

C(Y ) a unital ∗-homomorphism.<br />

Show that there is a continuous map f : Y → X such that ϕ(g) = g ◦ f, g ∈<br />

C(X). (Hint : For each y ∈ Y, g → ϕ(g)(y) defines a character of C(X). Use<br />

Exercise 1.83.)<br />

Exercise 1.90. Let X and Y be locally compact Hausdorff spaces. A continuous<br />

function f : Y → X is proper when the inverse image of every compact subset of X<br />

is compact in Y , i.e. when<br />

K ⊆ X compact ⇒ f −1 (K) ⊆ Y compact.<br />

a) Let f : Y → X be a continuous proper map. Show that there is a ∗homomorphism<br />

ψ : C0(X) → C0(Y ) such that ψ(h) = h ◦ f.<br />

b) Let ϕ : C0(X) → C0(Y ) be a ∗-homomorphism. Show that there is a proper<br />

continuous map f : Y → X such that ϕ(h) = h ◦ f for all h ∈ C0(X).<br />

c) Let ϕ and f be as in b). Show that ϕ is injective if and only if f is surjective.<br />

d) Let ϕ and f be as in b). Show that ϕ is isometric if and only if f is surjective.<br />

e) Show that C0(X) ≃ C0(Y ) if and only if there is a homeomorphism f : Y → X.<br />

Exercise 1.91. Let ϕ : A → B be a ∗-homomorphism between C ∗ -<strong>algebras</strong> A<br />

and B.<br />

a) Show that ϕ(a)) ≤ a when a ∈ A is selfadjoint. (Hint: One possibility is<br />

to reduce to the case when A and B are abelian, and then use Exercise ??.)<br />

b) Show that ϕ(a)) ≤ a for all a ∈ A.<br />

c) Show that ϕ is injective if and only if ϕ is isometric.<br />

Exercise 1.92. Let α be an automorphism of Mn (i.e. α is a ∗-isomorphism α :<br />

Mn → Mn). Show that there is a unitary U ∈ Mn such that α(A) = UAU ∗ , A ∈ Mn.


38 1. FUNDAMENTALS<br />

(Hint : Let {eij} be the standard set of matrix units in Mn. Show that α(e11)Mne11<br />

is spanned by a partial isometry v and consider<br />

n<br />

α(ej1)ve1j.)<br />

j=1<br />

Exercise 1.93. Let p, q ∈ Mn be projections. Show that there is a partial<br />

isometry v ∈ Mn such that vv ∗ = p and v ∗ v = q if and only if Tr(p) = Tr(q).<br />

Exercise 1.94. Let a be a selfadjoint element of a C ∗ -algebra A. Show that a<br />

is a projection if and only if σ(a) ⊆ {0, 1}.<br />

Exercise 1.95. Let X be a locally compact, non-compact Hausdorff space (e.g.<br />

X = R). Then A = C0(X) is an Abelian non-unital <strong>C∗</strong>-algebra and we can consider<br />

A ⊆ Â ⊆ M(A).<br />

1) Show that M(A) is ∗-isomorphic to the <strong>C∗</strong>-algebra Cb(X) of continuous<br />

bounded functions on X.<br />

2) Show that  is ∗-isomorphic to the <strong>C∗</strong>-algebra of continuous functions f :<br />

X → C which ’has a limit at infinity’, i.e. satisfies that there is a λ ∈ C<br />

such that<br />

∀ǫ > 0 ∃K ⊆ X compact : |f(x) − λ| ≤ ǫ ∀x ∈ X\K .<br />

3) Show that  ≃ C(Y ) where Y is a compact Hausdorff space with a point<br />

y0 ∈ Y such that X is homeomorphic to Y \{y0}. (Y is the one-point<br />

compactification of X.)<br />

4) Show that M(A) ≃ C(β(X)), where β(X) is the Stone-Cech-compactification<br />

of X.<br />

2. Inductive limits of C ∗ -<strong>algebras</strong><br />

2.1. The definition and the universal property. By a sequence of C ∗ -<br />

<strong>algebras</strong> we mean a countable family A1, A2, A3, . . . of C ∗ -<strong>algebras</strong> together with ∗<br />

-homomorphisms ϕn : An → An+1, n = 1, 2, 3, . . .. Such a sequence will sometimes<br />

be denoted<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

and sometimes more compressed as (An, ϕn). The ∗-homomorphisms ϕn will be<br />

called the connecting maps. If the An’s are unital C ∗ -<strong>algebras</strong> and the connecting<br />

maps all unital, we call the sequence a unital sequence of C ∗ - <strong>algebras</strong>. In any case,<br />

we will use the notation ϕm,n for the composition ϕm,n = ϕm−1 ◦ ϕm−2 ◦ · · · ◦ ϕn :<br />

An → Am when n < m. So we have that ϕn = ϕn+1,n. It is convenient to define<br />

ϕn,n to be the identity map on An.<br />

Assume now that we are given a sequence of C ∗ -<strong>algebras</strong> (An, ϕn). We will<br />

construct from this sequence a new C ∗ -algebra as follows. First we introduce an<br />

equivalence relation ≡ on the disjoint union<br />

∞<br />

n=1<br />

by writing a ≡ b when a ∈ An, b ∈ Am and there is a k ≥ max{n, m} such that<br />

ϕk,n(a) = ϕk,m(b). For a ∈ ∞<br />

n=1 An we let [a] denote the equivalence class containing<br />

An


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 39<br />

a. It is straightforward to check that we can make ∞<br />

n=1 An/ ≡ into a ∗-algebra by<br />

defining<br />

[a] + [b] = [ϕk,n(a) + ϕk,m(b)], a ∈ An, b ∈ Am, k ≥ max{n, m},<br />

[a][b] = [ϕk,n(a)ϕk,m(b)], a ∈ An, b ∈ Am, k ≥ max{n, m},<br />

λ[a] = [λa], a ∈ An, λ ∈ C,<br />

and<br />

[a] ∗ = [a ∗ ], a ∈ An.<br />

To get a <strong>C∗</strong>-algebra out of ∞ n=1 An/ ≡ we first introduce a semi-norm by<br />

[a] = lim<br />

k→∞ ϕk,n(a), a ∈ An.<br />

This limit exists because a ∗-homomorphism between <strong>C∗</strong>- <strong>algebras</strong> never increases<br />

the norm, so that the sequence ϕk,n(a), k ≥ n, is decreasing. It is<br />

straightforward to check that · defines a semi-norm on ∞ n=1 An/ ≡ satisfying<br />

xy ≤ xy, x ∗ = x, x ∗ x = x 2 , (2.1)<br />

for all x, y ∈ ∞ n=1 An/ ≡. In particular, it follows that<br />

∞<br />

I = {x ∈ An/ ≡ : x = 0}<br />

is a two-sided ∗-invariant ideal in ∞ n=1 An/ ≡. The quotient space<br />

∞<br />

( An/ ≡)/I<br />

is therefore a ∗-algebra and<br />

n=1<br />

n=1<br />

x + I = x, x ∈<br />

∞<br />

An/ ≡,<br />

defines norm on ( ∞ n=1 An/ ≡)/I. Let q : ∞ n=1 An/ ≡ → ( ∞ n=1 An/ ≡)/I denote the<br />

quotient map. Note that (2.1) also holds in ( ∞ n=1 An/ ≡)/I so that this ∗-algebra is<br />

almost a <strong>C∗</strong>-algebra, the only missing property is the completeness of the norm. We<br />

define the inductive limit of the sequence (An, ϕn) to be the C ∗ -algebra we obtain<br />

by completing ( ∞<br />

n=1 An/ ≡)/I in the norm · and we denote it by lim<br />

−→ (An, ϕn),<br />

or simply by A∞ when no ambiguity is possible.<br />

Let ι : ∞<br />

n=1 An/ ≡ → A∞ be the ∗-homomorphism obtained by composing the<br />

quotient map q with the canonical imbedding of ( ∞<br />

n=1 An/ ≡)/I into its completion,<br />

A∞. Define<br />

by<br />

n=1<br />

ϕ∞,n : An → A∞<br />

ϕ∞,n(a) = ι([a]), a ∈ An.<br />

These ∗-homomorphisms will be referred to as the canonical maps. It is clear<br />

from the construction that<br />

ϕ∞,n(An) ⊆ ϕ∞,n+1(An+1), n ∈ N, (2.2)


40 1. FUNDAMENTALS<br />

and that<br />

∞<br />

ϕ∞,n(An) = A∞. (2.3)<br />

n=1<br />

Furthermore, we shall frequently use the following properties. Let a ∈ An, b ∈ Am.<br />

Then<br />

and<br />

ϕ∞,n(a) = ϕ∞,m(b) ⇔ lim<br />

k→∞ ϕk,n(a) − ϕk,m(b) = 0 (2.4)<br />

ϕ∞,n(a) = lim<br />

k→∞ ϕk,n(a). (2.5)<br />

Before we give some examples to familiarize the reader with inductive limits, we<br />

will first emphasize the categorial properties of the construction. Assume that we<br />

are given a C ∗ -algebra B and ∗-homomorphisms ψn : An → B, n ∈ N, which are<br />

compatible with the connecting maps of the sequence (An, ϕn), in the sense that<br />

ψn+1 ◦ ϕn = ψn<br />

for all n ∈ N. We can then define a ∗-homomorphism<br />

such that<br />

ψ : A∞ → B<br />

ψ ◦ ϕ∞,n = ψn<br />

(2.6)<br />

(2.7)<br />

for all n ∈ N. Indeed, we first obtain a ∗-homomorphism ψ0 : ∞<br />

n=1 An/ ≡ → B<br />

by setting ψ0([a]) = ψn(a), a ∈ An. This is welldefined because of (2.6). Since<br />

ψ0([a]) = ψk(ϕk,n(a)) ≤ ϕk,n(a) for all k ≥ n, we see that ψ0([a]) ≤ [a].<br />

Thus ψ0 induces a ∗-homomorphism ψ00 : ( ∞ n=1 An/ ≡)/I → B such that ψ00 ◦ q =<br />

ψ0. Since ψ00(q(x)) = ψ0(x) ≤ x = q(x), x ∈ ∞ n=1 An/ ≡, we see that ψ00<br />

extends by continuity to a ∗-homomorphism ψ : A∞ → B which satisfies (2.7) by<br />

construction. Note that the condition (2.7) determines ψ uniquely because of (2.3).<br />

For later reference we state this important property of the construction as a<br />

lemma.<br />

Lemma 2.1. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of C ∗ -<strong>algebras</strong> and A∞ = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra. If ψn : An → B is a sequence of ∗-homomorphisms into the C ∗ -algebra<br />

B such that ψn+1 ◦ϕn = ψn, n ∈ N, then there is one and only one ∗-homomorphism<br />

ψ : A∞ → B such that ψ ◦ ϕ∞,n = ψn, n ∈ N.<br />

The property described in Lemma 2.1 is sometimes referred to as the universal<br />

property of the inductive limit C ∗ - algebra, although it is really a property of the<br />

family ϕ∞,n : An → A∞, n = 1, 2, 3, . . ., rather than of the C ∗ -algebra A∞ in itself.<br />

It is this universal property which characterizes the construction, see Exercis 2.18.<br />

Proposition 2.2. Let (An, ϕn) and (Bn, ψn) be two sequences of C ∗ -<strong>algebras</strong>.<br />

Assume that there are sequences k1 < k2 < k3 < . . . and m1 < m2 < m3 < . . . in N


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 41<br />

and ∗-homomorphisms µi : Aki → Bmi , λi : Bmi → Aki+1 such that ϕki+1,ki = λi ◦ µi<br />

and µi+1 ◦ λi = ψmi+1,mi for all i ∈ N, i.e. such that the infinite diagram<br />

ϕk2 ,k ϕk 1 3 ,k ϕk 2 4 ,k3 Ak1<br />

<br />

Ak2<br />

<br />

Ak3<br />

<br />

µ1<br />

<br />

<br />

λ1 <br />

<br />

<br />

<br />

µ2<br />

<br />

<br />

λ2 <br />

<br />

<br />

<br />

µ3<br />

<br />

· · ·<br />

<br />

λ3 <br />

<br />

<br />

<br />

<br />

· · ·<br />

Bm1<br />

<br />

Bm2<br />

<br />

Bm3<br />

ψm2 ,m1 ψm3 ,m2 ψm4 ,m3 commutes. It follows that lim(An,<br />

ϕn) ≃ lim(Bn,<br />

ψn). In fact, there is a ∗-isomorphism<br />

−→ −→<br />

µ : lim(An,<br />

ϕn) → lim(Bn,<br />

ψn) such that µ ◦ ϕ∞,kn = ψ∞,mn ◦ µn for all n ∈ N.<br />

−→ −→<br />

Proof. To simplify notation, set A∞ = lim(An,<br />

ϕn) and B∞ = lim(Bn,<br />

ψn). Let<br />

−→ −→<br />

i ∈ N and a ∈ Ai. Then the sequence<br />

where kj ≥ i, is constant, and we set<br />

ψ∞,mj ◦ µj ◦ ϕkj,i,<br />

µi = lim<br />

j→∞ ψ∞,mj ◦ µj ◦ ϕkj,i.<br />

This is a ∗-homomorphism µi : Ai → B∞ and µi+1 ◦ ϕi = limj→∞ ψ∞,mj ◦ µj ◦<br />

ϕkj,i+1 ◦ ϕi+1 = limj→∞ ψ∞,mj ◦ µj ◦ ϕkj,i = µi for all i. It follows therefore from<br />

the universal property of the inductive limit construction, Lemma 2.1 that there is<br />

a ∗-homomorphism µ : A∞ → B∞ such that µ ◦ ϕ∞,i = µi for all i. In particular,<br />

µ ◦ ϕ∞,kn ◦ ϕ∞,kn = limj→∞ ψ∞,mj ◦ µj ◦ ϕkj,kn = ψ∞,mn ◦ µn for all n ∈ N.<br />

It remains to show that µ is a ∗-isomorphism. To this end, let iN and b ∈ Bi.<br />

Then the sequence<br />

ϕ∞,kj+1 ◦ λj ◦ ψmj,i,<br />

where mj ≥ i, is constant, and we set<br />

λi = lim<br />

j→∞ ϕ∞,kj+1 ◦ λj ◦ ψmj,i.<br />

The λi’s are ∗-homomorphisms and λi+1 ◦ψi = λi for all i. By the universal property<br />

of the inductive limit construction, Lemma 2.1 that there is a ∗-homomorphism<br />

λ : B∞ → A∞ such that λ ◦ ψ∞,i = λi for all i. Note that<br />

λ ◦ µ ◦ ϕ∞,i = λ ◦ µi = λ ◦ lim<br />

j→∞ ψ∞,mj ◦ µj ◦ ϕkj,i<br />

= lim λ ◦ ψ∞,mj<br />

j→∞ ◦ µj<br />

λmj ◦ ϕkj,i = lim<br />

j→∞<br />

◦ µj ◦ ϕkj,i<br />

= lim lim ϕ∞,kl+1<br />

j→∞ l→∞ ◦ λl ◦ ψml,mj ◦ µj ◦ ϕkj,i<br />

= lim lim ϕ∞,kl+1 ◦ ϕkl+1,kj ◦ ϕkj,i<br />

j→∞ l→∞<br />

= ϕ∞,i.<br />

Since this holds for all i we see that λ ◦µ is the identity on A∞. A similar argument<br />

shows that µ ◦ λ is the identity on B∞.<br />

<br />

Example 2.3. The purpose of this relatively long example is to describe the C ∗ -<br />

algebra one gets as the inductive limit of a unital sequence of Abelian C ∗ -<strong>algebras</strong>.


42 1. FUNDAMENTALS<br />

So consider a sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

of unital Abelian C ∗ -<strong>algebras</strong> An with unital connecting maps. By Theorem 1.31<br />

there are compact Hausdorff spaces Xn and ∗-isomorphisms λn : An → C(Xn) for<br />

all n ∈ N. Partly for convenience we restrict the attention to the case where each<br />

Xn is metrizable.<br />

By Exercise 1.89 we know that there are continuous maps fn : Xn+1 → Xn such<br />

that the diagram<br />

λn<br />

An<br />

⏐<br />

<br />

ϕn<br />

−−−→ An+1<br />

C(Xn)<br />

commutes when ψn : C(Xn) → C(Xn+1) is the ∗-homomorphism induced by fn, viz.<br />

λn+1<br />

⏐<br />

<br />

ψn<br />

−−−→ C(Xn+1)<br />

ψn(g) = g ◦ fn, g ∈ C(Xn).<br />

By Proposition 2.2 we therefore have that lim(An,<br />

ϕn) ≃ lim(C(Xn),<br />

ψn). Now,<br />

−→ −→<br />

it is fairly clear that the inductive limit of a unital sequence of Abelian <strong>C∗</strong>-<strong>algebras</strong> must be unital and Abelian itself. Thus there must be a compact Hausdorff space<br />

X such that<br />

lim<br />

−→ (C(Xn), ψn) ≃ C(X).<br />

Also it must depend only on the Xn’s and the fn’s. Here it is. Set<br />

∞<br />

X = {(xi) ∈ Xi : xn = fn(xn+1) for all n ∈ N}.<br />

i=1<br />

X is clearly a closed subset of the infinite product ∞<br />

n=1 Xn and is therefore compact<br />

by Tychonoffs theorem. X is the inverse limit space (sometimes called the projective<br />

limit ) of the sequence<br />

X1<br />

<br />

X2<br />

<br />

X3<br />

X4<br />

<br />

f1 f2 f3<br />

. . .<br />

(by definition) and is usually denoted by lim(Xn,<br />

fn). Let us prove that lim(C(Xn),<br />

ψn) ≃<br />

←− −→<br />

C(X). Let pn : X → Xn be the map which projects to the n’th coordinate,<br />

i.e. pn (xi) = xn, and set Φn(g) = g ◦ pn, g ∈ C(Xn). Then the Φn’s are<br />

∗-homomorphisms Φn : C(Xn) → C(X) such that Φn+1 ◦ ψn = Φn. Indeed,<br />

Φn+1 ◦ ψn(g) = Φn+1(g ◦ fn) = g ◦ fn ◦ pn+1 = g ◦ pn bacause fn ◦ pn+1 = pn<br />

on X. By the universal property of the inductive limit, Lemma 2.1, there is a<br />

∗-homomorphism Φ : lim(C(Xn),<br />

ψn) → C(X) such that Φ ◦ ψ∞,n = Φn for all n.<br />

−→<br />

We first prove that Φ is an isometry. So let g ∈ C(Xn) for some n. It will suffice<br />

to show that<br />

Φ(ψ∞,n(g)) = Φn(g) = ψ∞,n(g).<br />

By (2.5)<br />

ψ∞,n(g) = lim<br />

k→∞ ψk,n(g).<br />

For each k > n choose xk ∈ Xk such that<br />

|g ◦ fn ◦ fn+1 ◦ · · · ◦ fk−1(xk)| = g ◦ fn ◦ fn+1 ◦ · · · ◦ fk−1 = ψk,n(g).<br />

f4


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 43<br />

By compactness of Xn there is a sequence m n 1 < mn 2 < mn 3<br />

sequence<br />

{fn ◦ fn+1 ◦ · · · ◦ fm n k −1(xm n k )}<br />

converges in Xn, say to yn. Next there is a subsequence m n+1<br />

1<br />

of mn 1 < mn2 < mn3 < . . . such that<br />

{fn+1 ◦ fn+2 ◦ · · · ◦ f n+1<br />

mk −1(xm n+1<br />

k<br />

)}<br />

< . . . in N such that the<br />

< mn+1<br />

2<br />

< mn+1<br />

3<br />

< . . .<br />

converges in Xn+1, say to yn+1. (It is here we use the assumption that Xn is<br />

metrizable.) Continuing in this way we get elements yj ∈ Xj, j ≥ n, such that<br />

yj = limk→∞ fj ◦fj+1 ◦· · ·◦f j<br />

m < . . . in N<br />

k−1(xm j ) for some sequence m<br />

k<br />

j<br />

1 < mj 2 < mj3<br />

and some subsequence m j+1<br />

1 < m j+1<br />

2 < m j+1<br />

3 < . . . of m j<br />

1 < m j<br />

2 < m j<br />

3 < . . . . Then<br />

fj(yj+1) = fj( lim fj+1 ◦ fj+2 ◦ · · · ◦ f j+1<br />

k→∞<br />

mk −1(xm j+1<br />

k<br />

lim<br />

k→∞ fj ◦ fj+1 ◦ · · · ◦ f j+1<br />

mk −1(xm j+1)<br />

= yj<br />

k<br />

)) =<br />

for all j ≥ n so if we set yi = fi ◦ fi+1 ◦ · · · ◦ fn−1(yn), i < n, then the sequence<br />

y = (yi) is in X. Note that<br />

Φn(g) = g ◦ pn ≥ |g ◦ pn(y)| = |g(yn)| = lim |g(fn ◦ fn+1 ◦ · · · ◦ fm<br />

k→∞ n k −1(xmn))| = k<br />

lim<br />

k→∞ ψmn k ,n(g) = ψ∞,n(g).<br />

Thus<br />

Φ(ψ∞,n(g)) ≥ ψ∞,n(g)<br />

and the reversed inequality is automatic because Φ is a ∗-homomorphism between<br />

<strong>C∗</strong>-<strong>algebras</strong>. We have shown that Φ is an isometric ∗-homomorphism. It is clear<br />

that Φ is unital (Φ(ψ∞,n(1)) = 1), so to conclude that Φ is surjective and hence<br />

is a ∗-isomorphism, it suffices by the Stone-Weierstrass theorem to check that<br />

Φ(lim(C(Xn),<br />

ψn)) separates the points of X. So consider (xi) = (zi) in X. There<br />

−→<br />

is an n ∈ N such that xn = zn and hence there is an element g ∈ C(Xn) such that<br />

g(xn) = g(zn). Since<br />

and similarly<br />

Φ(ψ∞,n(g)) (xi) = Φn(g) (xi) = g(xn),<br />

Φ(ψ∞,n(g)) (zi) = Φn(g) (zi) = g(zn),<br />

this completes the proof.<br />

Speaking in the language of category theory, the conclusion that<br />

lim<br />

−→ (C(Xn), ψn) ≃ C(lim(Xn,<br />

fn)),<br />

←−<br />

tells us that the contravariant functor X → C(X) from compact Hausdorff spaces<br />

to Abelian unital <strong>C∗</strong>-<strong>algebras</strong> is continuous.<br />

2.2. AF-<strong>algebras</strong>. We now come to an exciting and surprisingly rich class of<br />

C ∗ -<strong>algebras</strong> whose structure is very well understood and for which there is a neat<br />

classification, up to isomorphism, by fairly simple invariants.<br />

Definition 2.4. A C ∗ -algebra A is called approximately finite dimensional (AFalgebra,<br />

for short) when there is an increasing sequence<br />

A1 ⊆ A2 ⊆ A3 ⊆ . . .


44 1. FUNDAMENTALS<br />

of finite dimensional <strong>C∗</strong>-sub<strong>algebras</strong> in A whose union is dense, i.e. satisfies that<br />

∞<br />

An = A.<br />

n=1<br />

We know from Theorem 1.60 that a finite dimensional C ∗ - algebra is ∗-isomorphic<br />

to a finite direct sum of matrix <strong>algebras</strong> :<br />

Mn1 ⊕ Mn2 ⊕ Mn3 ⊕ · · · ⊕ MnN<br />

The following lemma can therefore hardly come as a surprise.<br />

Lemma 2.5. A C ∗ -algebra A is an AF-algebra if and only if there is a sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

of finite direct sums of matrix <strong>algebras</strong> such that A ≃ lim<br />

−→ (An, ϕn).<br />

Proof. If F1 ⊆ F2 ⊆ F3 ⊆ . . . is a sequence of finite dimensional C ∗ -sub<strong>algebras</strong><br />

in A such that A = ∞<br />

n=1 Fn, we can use Theorem 1.60 to get finite direct sums of<br />

matrix <strong>algebras</strong> An and ∗- isomorphisms ψn : Fn → An, n ∈ N. Let in denote the<br />

inclusion Fn → Fn+1. Then ϕn = ψn+1 ◦ in ◦ ψ −1<br />

n : An → An+1 makes the diagram<br />

ψn<br />

Fn<br />

⏐<br />

<br />

An<br />

in<br />

−−−→ Fn+1<br />

ψn+1<br />

⏐<br />

<br />

ϕn<br />

−−−→ An+1<br />

commute and we can define ψ : ∞<br />

n=1 Fn → A∞ by ψ|Fn = ϕ∞,n ◦ ψn. Note that<br />

ψ(a) = limk→∞ ϕk,n ◦ ψn(a) = a because each ϕk,n and ψn is injective and<br />

hence isometric by Exercise 1.91. Thus ψ extends to an injective (isometric) ∗-<br />

homomorphism ψ : A → A∞ = lim<br />

−→ (An, ϕn). Obviously ϕ∞,n(An) = ψ(Fn) is in the<br />

image of ψ for all n and hence ψ must be surjective.<br />

Conversely, if A ≃ A∞ = lim<br />

−→ (An, ϕn) for a sequence of finite direct sums of<br />

matrix <strong>algebras</strong>, let ψ : A∞ → A be a ∗-isomorphism. Then Fn = ψ(ϕ∞,n(An)) is a<br />

finite dimensional C ∗ - subalgebra of A, Fn ⊆ Fn+1 for all n and ∞<br />

n=1 Fn = A. <br />

The AF-<strong>algebras</strong> are thus just the C ∗ -<strong>algebras</strong> which arise as an inductive limit<br />

of a sequence of finite direct sums of matrix <strong>algebras</strong>. The structure of matrix<br />

<strong>algebras</strong>, or a finite direct sum of them, is as uncomplicated as one could wish<br />

and hence the problem, which must be overcome in order to gain a more detailed<br />

insight into the nature of AF-<strong>algebras</strong>, is to determine how such finite dimensional<br />

C ∗ -<strong>algebras</strong> can be put together to form an inductive limit. Thus the first task is to<br />

obtain a satisfying description of the ∗- homomorphisms between finite direct sums<br />

of matrix <strong>algebras</strong>.<br />

Let A = Mn1 ⊕Mn2 ⊕· · ·⊕MnN and B = Mm1 ⊕Mm2 ⊕· · ·⊕MmM . Let S = (sij)<br />

be an M × N matrix with entries from N such that<br />

N<br />

sijnj ≤ mi, i = 1, 2, . . ., M. (2.8)<br />

j=1<br />

We can then define a ∗-homomorphism ϕS : A → B such that<br />

ϕS(a1, a2, . . .,aN) = (ϕ1(a1, a2, . . .,aN), ϕ2(a1, a2, . . .,aN), . . ....,ϕM(a1, a2, . . ., aN))


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 45<br />

for all (a1, a2, . . .,aN) ∈ A, where ϕi : A → Mmi is the ∗-homomorphism which<br />

sends (a1, a2, . . .,aN) ∈ A to the block diagonal matrix<br />

diag <br />

a1, a1, . . .,a1 , a2, a2, . . .,a2 , . . ....,aN, aN, . . ., aN,<br />

0, 0, . . ., 0<br />

<br />

si1 times si2 times<br />

siN times r times<br />

where r = mi − N<br />

j=1 sijnj. Such a ∗-homomorphism will be called a standard<br />

homomorphism.<br />

For example, if A = M2 ⊕ M3 and B = M7, the 1 × 2 matrix S = (s11, s12) =<br />

(2, 1) determines the standard homomorphism ϕS : M2 ⊕ M3 → M7 which sends<br />

(a, b) = ((aij), (bij)) ∈ M2 ⊕ M3 to<br />

diag ⎛<br />

a11 a12 0 0 0 0<br />

⎞<br />

0<br />

⎜a21<br />

⎜ 0<br />

⎜<br />

a, a, b) = ⎜ 0<br />

⎜ 0<br />

⎝<br />

0<br />

a22<br />

0<br />

0<br />

0<br />

0<br />

0<br />

a11<br />

a21<br />

0<br />

0<br />

0<br />

a12<br />

a22<br />

0<br />

0<br />

0<br />

0<br />

0<br />

b11<br />

b21<br />

0<br />

0<br />

0<br />

b12<br />

b22<br />

0 ⎟<br />

0 ⎟<br />

0 ⎟.<br />

⎟<br />

b13⎟<br />

⎠<br />

b23<br />

0 0 0 0 b31 b32 b33<br />

Note that the matrix S = (sij) is easy to recover from ϕS; if {e d ij : i, j =<br />

1, 2, . . ., nd, d = 1, 2, . . ., N} is the standard system of matrix units in A, then<br />

sij = Tri(ϕS(e j<br />

11))<br />

where Tri : B → C is the composition of the projection B → Mmi onto the i’th<br />

component of B with the trace Tr of Mmi , i.e.<br />

Tri(b1, b2, . . .,bM) = Tr(bi), (b1, b2, . . .,bM) ∈ B.<br />

A very useful way to represent a standard homomorphism is by using a diagram.<br />

The diagram, called a Bratteli diagram, corresponding to the standard homomorphism<br />

given by the matrix S = (sij) is obtained as follows. First draw a horizontal<br />

row of N vertices, labelled n1, n2, . . ., nN and below that another horizontal row of<br />

M vertices labelled m1, m2, . . .,mM. Then draw sij edges from the vertex labelled nj<br />

to the vertex labelled mi for all i, j. For example the standard map M2 ⊕M3 → M7<br />

described above has Bratteli diagram<br />

2 <br />

3<br />

<br />

<br />

<br />

<br />

<br />

<br />

7<br />

The importance of the standard homomorphisms comes from the following two<br />

results.<br />

Lemma 2.6. Let A = Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN , B = Mm1 ⊕ Mm2 ⊕ · · · ⊕ MmM<br />

and let ψ : A → B be a ∗-homomorphism.<br />

There is then a unitary u ∈ B and a standard homomorphism ϕ : A → B such<br />

that<br />

uψ(a)u ∗ = ϕ(a), a ∈ A.


46 1. FUNDAMENTALS<br />

Proof. Let {e d ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N} be the standard system of<br />

matrix units in A. Set sij = Tri(ψ(e j<br />

11)), j = 1, 2, . . ., N, i = 1, 2, . . ., M. Note that<br />

Tri(ψ(e j<br />

j<br />

kk )) = Tri(ψ(ek1 )ψ(ej<br />

j<br />

1k )) = Tri(ψ(e1k )ψ(ej<br />

j<br />

k1 )) = Tri(ψ(e11 ))<br />

for all k = 1, 2, . . ., nj. Thus sijnj = nj j<br />

k=1 Tri(ψ(ekk )) and<br />

N<br />

j=1<br />

sijnj =<br />

N<br />

nj <br />

j=1 k=1<br />

Tri(ψ(e j<br />

kk<br />

)) = Tri(ψ( <br />

j,k<br />

e j<br />

kk )) = Tri(ψ(1)) ≤ Tri(1) = mi.<br />

Thus (sij) satisfies (2.8) and hence defines a standard homomorphism ϕ : A → B.<br />

Let πi : B → Mmi be the projection and set ψi = πi ◦ ψ, ϕi = πi ◦ ϕ. It will suffice<br />

to construct, for each i, a unitary ui ∈ Mmi such that uiψi(a)u∗ i = ϕi(a), a ∈ A,<br />

because then u = (u1, u2, . . .,uM) ∈ B will do the job. So fix i ∈ {1, 2, . . ., M}.<br />

Since<br />

Tr(ψi(e j<br />

11)) = Tr(ϕi(e j<br />

11)),<br />

we know from Exercise 1.93, that there is a partial isometry vj ∈ Mmi such that<br />

vjv∗ j<br />

j = ϕi(e11) and v∗ jvj = ψi(e j<br />

11). Set v = N nj j j<br />

j=1 k=1 ϕi(ek1 )vjψi(e1k ) and note<br />

that<br />

vv ∗ = <br />

<br />

<br />

) ) =<br />

<br />

j,k<br />

j,k<br />

ϕi(e j j<br />

k1 )vjψi(e1k ϕi(e j j<br />

k1 )ϕi(e11)ϕi(e j<br />

1k<br />

s,t<br />

) = <br />

j,k<br />

ψi(e t s1 )v∗ t ϕi(e t 1s<br />

ϕi(e j<br />

kk ) = ϕi(1),<br />

where we have used that vjψi(e j<br />

11 )v∗ j = vjv ∗ j vjv ∗ j<br />

Since<br />

j,k<br />

ϕi(e j j<br />

k1 )vjψi(e11 )v∗ j<br />

jϕi(e1k ) =<br />

= ϕi(e j<br />

11 ). Similarly, v∗ v = ψi(1).<br />

Tr(1 − ψi(1)) = mi − Tr(ψi(1)) = mi − Tr(v ∗ v) = mi − Tr(vv ∗ ) =<br />

mi − Tr(ϕi(1)) = Tr(1 − ϕi(1)),<br />

we can use Exercise 1.93 again to get a partial isometry w ∈ Mmi such that ww∗ =<br />

1 −ϕi(1) and w∗w = 1 −ψi(1). Then ui = v +w is a unitary in Mmi (CHECK!) and<br />

uiψi(e d kl )u∗ i = vψi(e d kl )v∗ = <br />

b,c,s,t<br />

ϕi(e b c1 )vbψi(e b 1c )ψi(e d kl )ψi(e s t1 )v∗ sϕi(e s 1t ) =<br />

ϕi(e d k1 )vdψi(e d 11 )v∗ d ϕi(e d 1l ) = ϕi(e d k1 )ϕi(e d 11 )ϕi(e d 1l ) = ϕi(e d kl ).<br />

Since this holds for all k, l = 1, 2, . . ., nd, d = 1, 2, . . ., N, we see that uiψi(a)u ∗ i =<br />

ϕi(a), a ∈ A.<br />

<br />

Theorem 2.7. Let A be an AF-algebra. There is then a sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

of finite direct sums of matrix <strong>algebras</strong> with connecting maps that are all standard<br />

homomorphisms such that A ≃ lim<br />

−→ (An, ϕn).<br />

Proof. By Lemma 2.5 there is a sequence (An, ψn) of finite direct sums of<br />

matrix <strong>algebras</strong> such that A ≃ lim(An,<br />

ψn). So the only problem we have is that<br />

−→<br />

we must substitute the ψn’s by standard homomorphisms. By Proposition 2.2 it


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 47<br />

is enough to find standard homomorphisms ϕn : An → An+1 and ∗-automorphisms<br />

αn : An → An such that the diagram<br />

α1<br />

ψ1 ψ2<br />

A1 −−−→ A2 −−−→ A3<br />

⏐<br />

<br />

α2<br />

⏐<br />

<br />

α3<br />

⏐<br />

<br />

ψ3<br />

−−−→ . . .<br />

ϕ1 ϕ2 ϕ3<br />

A1 −−−→ A2 −−−→ A3 −−−→ . . .<br />

commutes. This is done by induction using Lemma 2.6. The induction is started by<br />

taking α1 to be the identity map on A1. So assume that we have constructed the<br />

diagram up to<br />

αn−1<br />

An−1<br />

⏐<br />

<br />

An−1<br />

ψn−1<br />

−−−→ An<br />

αn<br />

⏐<br />

<br />

ϕn−1<br />

−−−→ An<br />

Then ψn ◦ α −1<br />

n : An → An+1 is unitarily conjugate to a standard homomorphism ϕn<br />

by Lemma 2.6; i.e. there is a unitary v ∈ An+1 such that ϕn = Ad v ◦ ψn ◦ α −1<br />

n is a<br />

standard homomorphism. If we set αn+1 = Ad v, the diagram<br />

αn<br />

An<br />

⏐<br />

<br />

An<br />

ψn<br />

−−−→ An+1<br />

αn+1<br />

⏐<br />

<br />

ϕn<br />

−−−→ An+1<br />

commutes and hence we have completed the induction step. <br />

We now consider the unital case. Note that a standard homomorphism ϕ :<br />

Mn1 ⊕ · · · ⊕MnN → Mm1 ⊕ · · · ⊕MmM given by the matrix (sij) is unital if and only<br />

if<br />

N<br />

sijnj = mi, i = 1, 2, . . ., M. (2.9)<br />

j=1<br />

As one would expect, a unital AF-algebra can be realized as the inductive limit<br />

of a sequence of finite direct sums of matrix <strong>algebras</strong> with unital standard homomorphisms<br />

as connecting maps. The starting point for the proof of this is first<br />

to write the given unital AF-algebra A as the closure of an increasing sequence<br />

A1 ⊆ A2 ⊆ A3 ⊆ . . . of finite dimensional <strong>C∗</strong>-sub<strong>algebras</strong>, viz.<br />

A =<br />

∞<br />

An.<br />

n=1<br />

By Lemma 1.61 each An contains a unit for it self, say pn ∈ An. Then limn→∞ pna =<br />

a for all a ∈ A. Indeed, if a ∈ A and ǫ > 0 are given, there is an N ∈ N and b ∈ AN<br />

. Then<br />

such that a − b < ǫ<br />

2<br />

pna − a ≤ pna − pnb + pnb − b + b − a = pn(a − b) + b − a < ǫ<br />

for all n ≥ N. In particular we see that<br />

lim<br />

n→∞ 1 − pn = lim pn1 − 1 = 0.<br />

n→∞<br />

So for all large enough n, 1−pn < 1<br />

2 . But 1−pn is a projection, so σ(1−pn) ⊆ {0, 1}<br />

(this follows from Exercise 1.93), and hence 1 − pn < 1 ⇔ pn = 1. So we see that


48 1. FUNDAMENTALS<br />

pn = 1 for all large enough n, say for n ≥ N. Thus A = <br />

n≥N An and 1 ∈ An for all<br />

n ≥ N. Consequently we might as well assume to begin with that 1 ∈ An for all n.<br />

It is now straightforward to check that with this assumption as point of departure,<br />

all the maps constructed in and betweeen Lemma 2.5 and Theorem 2.7 become<br />

unital. So we arrive at the following conclusion.<br />

Theorem 2.8. Let A be a unital AF-algebra. There is then a sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

of finite direct sums of matrix <strong>algebras</strong> with unital standard homomorphisms as connecting<br />

maps such that A ≃ lim<br />

−→ (An, ϕn).<br />

Thanks to Theorem 2.7 and Theorem 2.8, an AF-algebra A can be described by<br />

use of infinite Bratteli diagrams obtained as follows. Write<br />

A ≃ lim<br />

−→ (An, ϕn)<br />

for some sequence (An, ϕn) of finite direct sums of matrix <strong>algebras</strong> with standard<br />

homomorphisms as connecting maps. The Bratteli diagram for A is then the infinite<br />

graph obtained by drawing the Bratteli diagrams of the standard homomorphisms<br />

ϕn : An → An+1 below one another. For example<br />

2 2<br />

<br />

<br />

<br />

4 4<br />

<br />

<br />

<br />

8 8<br />

<br />

<br />

<br />

16 <br />

16<br />

<br />

<br />

<br />

<br />

. . .<br />

is the Bratteli diagram for the AF-algebra which is the inductive limit of the sequence<br />

M2 ⊕ M2<br />

ϕ1 <br />

M4 ⊕ M4<br />

ϕ2 <br />

M8 ⊕ M8<br />

ϕ3 <br />

. . .<br />

where ϕn : M2n ⊕ M2n → M2n+1 ⊕ M2n+1 is the standard homomorphism given by<br />

the matrix <br />

2 0<br />

1 1<br />

for all n ∈ N.<br />

Note that the AF-algebra we get is unital because the connecting maps are. This<br />

can be used to simplify the Bratteli diagram a little. Because, if it is stated that the<br />

connecting maps are unital, then the numbers occuring in row 2 and onwards are<br />

determined completely by the numbers in the first row and the diagram itself, by<br />

repeated use of (2.9). And we can always assume, in the unital case, that A1 = C so<br />

that the first row of vertices only contains one vertex which is labelled 1. Since this


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 49<br />

is a standard convention, one doesn’t even write that first label in. So usually, in the<br />

litterature, the AF-algebra above is defined as the unital AF-algebra with diagram<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

. . .<br />

In the non-unital case, however, it is important to indicate the integer labels<br />

on the vertices, because otherwise the matrix sizes involved in the sequence of <strong>C∗</strong>- <strong>algebras</strong> can not be determined from the diagram.<br />

We will now investigate an important special class of unital AF-<strong>algebras</strong>, namely<br />

the socalled UHF-<strong>algebras</strong>. These are the AF-<strong>algebras</strong> which we obtain as inductive<br />

limits of sequences<br />

Mn1<br />

ϕ1 <br />

Mn2<br />

ϕ2 <br />

Mn3<br />

ϕ3 <br />

. . .<br />

with unital connecting maps. It should be clear at this point that a unital <strong>C∗</strong>-algebra A is (isomorphic to) a UHF-algebra if and only if there is a sequence 1 ∈ A1 ⊆ A2 ⊆<br />

A3 ⊆ . . . of <strong>C∗</strong>-sub<strong>algebras</strong> in A such that each An is isomorphic to a matrix algebra<br />

and ∞ n=1 An = A.<br />

The UHF-<strong>algebras</strong> can be classified (without use of K-theory) in the following<br />

way. Let A be a UHF-algebra. Let P denote the set of prime numbers, remembering<br />

that 1 is not a prime. For each prime number p ∈ P we set<br />

FA(p) = sup{n ∈ N ∪ {0} : there is a unital ∗-homomorphism Mpn → A}.<br />

Then FA : P → N ∪ {∞}. The classification of the UHF-<strong>algebras</strong> can then be stated<br />

as follows.<br />

Theorem 2.9. a) Two UHF-<strong>algebras</strong> A and B are isomorphic if and only if<br />

FA = FB.<br />

b) For every function F : P → N ∪ {∞} there is a UHF-algebra A such that<br />

F = FA.<br />

For the proof of this theorem we need some lemmas. For applications later on,<br />

they are formulated and proved in a setting more general than what is required for<br />

the present applications.<br />

Lemma 2.10. Let a = a∗ be a selfadjoint element of the <strong>C∗</strong>-algebra A. If a2 −<br />

, then there is a projection p ∈ A such that a − p ≤ 2δ.<br />

a ≤ δ < 1<br />

4


50 1. FUNDAMENTALS<br />

Proof. Since σ(a 2 − a) = {t 2 − t : t ∈ σ(a)} ⊆ [−δ, δ] ⊆ [− 1<br />

4<br />

1 , ] an elementary<br />

4<br />

investigation of the function t ↦→ t2 −t shows that σ(a) ⊆ [−2δ, 2δ] ∪[1 −2δ, 1+2δ].<br />

Thus the characteristic function 1 1<br />

[ ,2] for the interval [1,<br />

2] is continuous on σ(a).<br />

2 2<br />

Hence p = 1 1<br />

[ 2 ,2](a) is a projection in A and since |1 [ 1 ,2](t) − t| ≤ 2δ, t ∈ σ(a), we<br />

2<br />

see that p − a ≤ 2δ.<br />

<br />

Lemma 2.11. When p, q are projections in the <strong>C∗</strong>-algebra A, x ∈ A, x ≤ 1,<br />

and<br />

xx ∗ − p ≤ δ < 1<br />

3 , x∗x − q ≤ δ < 1<br />

3<br />

then there is a partial isometry v ∈ A such that vv∗ = p, v∗v = q and x−v ≤ 4 √ δ.<br />

Proof. Set a = pxq and note that<br />

aa ∗ − p = pxqx ∗ p − p ≤ px(q − x ∗ x)x ∗ p + pxx ∗ xx ∗ p − p<br />

≤ δ + pxx ∗ xx ∗ p − p ≤ δ + p(xx ∗ − p)xx ∗ p + p 2 xx ∗ p − p<br />

≤ 2δ + pxx ∗ p − p = 2δ + p(xx ∗ − p)p ≤ 3δ < 1 .<br />

Similarly, a∗a − q ≤ 3δ < 1. Hence, by Lemma 1.10, aa∗ is invertible in the<br />

unital <strong>C∗</strong>-algebra pAp and a∗a is invertible in the unital <strong>C∗</strong>-algebra qAq. We<br />

set v = (aa∗ ) −1 2a where the functional calculus is done within pAp. Then vv∗ =<br />

(aa∗ ) −1 2aa∗ (aa∗ ) −1<br />

2 = p. In particular, v∗v is a projection by Lemma 1.55. Note<br />

that v ∗ v ∈ qAq so that qv ∗ v = v ∗ v. On the other hand v ∗ v = a ∗ (aa ∗ ) −1 a and<br />

v ∗ va ∗ a = a ∗ (aa ∗ ) −1 aa ∗ a = a ∗ pa = a ∗ a. In combination these identities show<br />

that (q − v ∗ v)a ∗ a = 0. Since a ∗ a is invertible in qAq, this implies that q = v ∗ v.<br />

To estimate x − v note first that<br />

v − a 2 = (aa ∗ ) −1 2a − a 2 = ((aa ∗ ) −1<br />

2 − p)a 2<br />

= [(aa ∗ ) −1<br />

2 − p]aa ∗ [(aa ∗ ) −1<br />

2 − p] = sup<br />

t∈σ(aa ∗ )<br />

≤ sup<br />

t∈σ(aa ∗ )<br />

= p − aa ∗ ≤ 3δ .<br />

Furthermore,<br />

|1 − 2 √ t + t|<br />

|1 − t| (since 0 ≤ 1 − 2 √ t + t ≤ 1 − t on σ(aa ∗ ) ⊆ [0, 1])<br />

x−px 2 = (1−p)xx ∗ (1−p) ≤ (1−p)xx ∗ (1−p)−(1−p)p(1−p)+(1−p)p(1−p) ≤ δ ,<br />

and similarly, x − xq 2 = (1 − q)x ∗ x(1 − q) ≤ δ. Hence<br />

x − a = x − pxq ≤ x − px + p(x − xq) ≤ 2 √ δ .<br />

In total, v − x ≤ v − a + a − x ≤ √ 3δ + 2 √ δ ≤ 4 √ δ. <br />

Lemma 2.12. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . . (2.10)<br />

be a sequence of C ∗ -<strong>algebras</strong> and A∞ = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra.<br />

If p1, p2, . . .,pN are mutually orthogonal projections in A∞ and ǫ > 0, then there<br />

is an m ∈ N and orthogonal projections q1, q2, . . ., qN ∈ Am such that ϕ∞,m(qi) −<br />

pi < ǫ, i = 1, 2, . . ., N.


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 51<br />

If the sequence (2.10) is unital and N<br />

i=1 pi = 1, then the qi’s can be chosen such<br />

that, in addition, N<br />

i=1 qi = 1.<br />

Proof. We use induction in N. The argument for the induction start, N = 1,<br />

is contained in the proof of the induction step and hence we only give the latter. So<br />

assume that the assertion holds for N − 1 projections and consider N orthogonal<br />

projections p1, p2, . . ., pN in A∞. Choose δ > 0 such that 22δ < 1<br />

4<br />

and 48δ <<br />

ǫ. There is then, by induction hypothesis, a k ∈ N and orthogonal projections<br />

q1, q2, . . .,qN−1 ∈ Ak such that<br />

ϕ∞,k(qi) − pi < δ<br />

, i = 1, 2, . . ., N − 1.<br />

N<br />

Choose n ≥ k and a ∈ An such that<br />

ϕ∞,n(a) − pN < δ<br />

N .<br />

By substituting 1<br />

2 (a + a∗ ) for a we may assume that a is selfadjoint. Note that<br />

lim<br />

k→∞ ϕk,n(a) < pN + δ ≤ 1 + δ<br />

so upon substituting a ∈ An by ϕl,n(a) ∈ Al for some l ≥ n, we can assume that<br />

a ≤ 1 + δ ≤ 2.<br />

Set Q = q1 + q2 + · · · + qN−1 ∈ Ak. Note that<br />

so that<br />

Furthermore,<br />

ϕ∞,k(Q)ϕ∞,n(a) = ϕ∞,n(a)ϕ∞,k(Q) < δ<br />

ϕ∞,k(Q)ϕ∞,n(a) + ϕ∞,n(a)ϕ∞,k(Q) − ϕ∞,k(Q)ϕ∞,n(a)ϕ∞,k(Q) < 3δ.<br />

ϕ∞,n(a) 2 − ϕ∞,n(a) ≤<br />

ϕ∞,n(a) 2 − ϕ∞,n(a)pN + ϕ∞,n(a)pN − pN + ϕ∞,n(a) − pN < 4δ.<br />

Choose m ≥ n such that<br />

and<br />

Set<br />

ϕm,k(Q)ϕm,n(a) + ϕm,n(a)ϕm,k(Q) − ϕm,k(Q)ϕm,n(a)ϕm,k(Q) < 3δ.<br />

ϕm,n(a) 2 − ϕm,n(a) < 4δ.<br />

b = ϕm,n(a) − ϕm,k(Q)ϕm,n(a) − ϕm,n(a)ϕm,k(Q) + ϕm,k(Q)ϕm,n(a)ϕm,k(Q)<br />

and note that b ≤ 2 + 3δ ≤ 3. Furthermore,<br />

b 2 − b ≤ 18δ + ϕ∞,n(a) 2 − ϕ∞,n(a) < 22δ,<br />

ϕ∞,m(b) − pN < 4δ,<br />

(2.11)<br />

and b ∈ {x ∈ Am : ϕm,k(Q)x = xϕm,k(Q) = 0}. The latter set is a C ∗ -subalgebra<br />

of Am so we can use Lemma 2.10 to conclude that there is a projection rN ∈ Am<br />

such that b − rN ≤ 44δ and ϕm,k(Q)rN = rNϕm,k(Q) = 0. Set ri = ϕm,k(qi), i =<br />

1, 2, . . ., N −1. Then r1, r2, . . ., rN are orthogonal projections in Am and ϕ∞,m(ri)−<br />

pi < 48δ < ǫ for all i = 1, 2, . . ., N. This completes the proof of the induction step<br />

and hence the proof in the non-unital case.


52 1. FUNDAMENTALS<br />

When (2.10) is unital and N<br />

i=1 pi = 1 we first conclude from the non-unital<br />

case that there is an k ∈ N and orthogonal projections r1, r2, . . .,rN ∈ Ak such that<br />

}. Then<br />

pi − ϕ∞,k(ri) < min{ǫ, 1<br />

N+1<br />

so for some m ≥ k<br />

1 − ϕ∞,k(<br />

N<br />

i=1<br />

ϕm,k(1 −<br />

ri) < N<br />

N + 1<br />

N<br />

ri) < 1<br />

i=1<br />

which implies that 1 = N<br />

i=1 ϕm,k(ri). Then qi = ϕm,k(ri), i = 1, 2, . . ., N, will have<br />

the desired properties. <br />

Proposition 2.13. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . . (2.12)<br />

be a sequence of C ∗ -<strong>algebras</strong> and A∞ = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra. If {e d ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N} is a set of matrix units in A∞<br />

and ǫ > 0, then there is an M ∈ N and a set {f d ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N}<br />

of matrix units in AM such that<br />

ϕ∞,M(f d ij ) − edij < ǫ<br />

for all i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N.<br />

If (2.12) is a unital sequence of <strong>C∗</strong>-<strong>algebras</strong> and N can be chosen such that N nd d=1 i=1 fd ii = 1.<br />

d=1<br />

nd<br />

i=1 ed ii = 1 then {f d ij}<br />

Proof. We first choose δ > 0 so small that 5δ < 1<br />

3 and 24√ δ + 4δ < ǫ.<br />

By Lemma 2.12 there is an m ∈ N and orthogonal projections q d ii ∈ Am, i =<br />

1, 2, . . ., nd, d = 1, 2, . . ., N, such that<br />

ϕ∞,m(q d ii ) − edii < δ<br />

for all d, i (and <br />

i,d qd ii = 1 when (2.12) is a unital sequence). For each d, j, choose<br />

wd j ∈ Ak, k ≥ m, such that<br />

ϕ∞,k(w d j ) − ed1j < δ.<br />

Since e d 1j ≤ 1 and limr→∞ ϕr,k(w d j) < e d 1j + δ ≤ 1 + δ, there is an r ≥ k such<br />

that<br />

ϕr,k(w d j) ≤ 1 + δ<br />

for all d, j. Set v d j = (1 + δ)−1 ϕr,k(w d j ) ∈ Ar and note that<br />

for all d, j. It follows that<br />

and<br />

for all d, j. Thus<br />

ϕ∞,r(v d j ) − ed 1j < (1 − (1 + δ)−1 )(1 + δ) + δ = 2δ<br />

ϕ∞,r(v d j)ϕ∞,r(v d j) ∗ − ϕ∞,m(q d 11) < 4δ + δ<br />

ϕ∞,r(v d j )∗ϕ∞,r(v d j ) − ϕ∞,m(q d jj ) < 4δ + δ


and<br />

2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 53<br />

ϕM,r(v d j )ϕM,r(v d j )∗ − ϕM,m(q d 11 ) ≤ 5δ<br />

ϕM,r(v d j )∗ϕM,r(v d j ) − ϕM,m(q d jj ) ≤ 5δ<br />

for some M ≥ r and all d, j. By Lemma 2.11 there is therefore a partial isometry<br />

sd j ∈ AM such that sd jsd ∗<br />

j = ϕM,m(qd 11), sd∗ d<br />

j sj = ϕM,m(qd jj) and sd j − ϕM,r(vd j ) ≤<br />

4 √ 5δ ≤ 12 √ δ. It follows that<br />

f d ij = sd ∗ d<br />

i sj , i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N,<br />

is a set of matrix units in AM (with N nd d=1 i=1 fd ii = 1 when (2.12) is a unital<br />

sequence) such that<br />

ϕ∞,M(f d ij ) − ed ij ≤ 24√ δ + 4δ < ǫ<br />

for all d, i, j. <br />

Corollary 2.14. Let<br />

Mn1<br />

ϕ1 <br />

Mn2<br />

ϕ2 <br />

Mn3<br />

ϕ3 <br />

. . .<br />

be a unital sequence of matrix <strong>algebras</strong> and A∞ = lim<br />

−→ (Mni , ϕi) the corresponding<br />

UHF-algebra. For any k ∈ N, there is a unital ∗-homomorphism ψ : Mk → A∞ if<br />

and only if k|ni for some i ∈ N.<br />

Proof. If k|ni there is a unital ∗-homomorphism µ : Mk → Mni , e.g. the<br />

standard homomorphism given by the 1×1 matrix ( ni<br />

k ). Then ψ = ϕ∞,ni ◦µ : Mk →<br />

A∞ is a unital ∗-homomorphism.<br />

Conversely, if ψ : Mk → A∞ is a unital ∗-homomorphism, let {eij} be the<br />

standard system of matrix units in Mk. Then {ψ(eij)} is a system of matrix units<br />

in A∞ with k<br />

i=1 ψ(eii) = 1 and by Proposition 2.13 there is a set {fij : i, j =<br />

1, 2, . . ., k} of matrix units in Mni for some i, such that k<br />

i=1 fii = 1 in Mni<br />

. We<br />

therefore obtain a unital ∗-homomorphism µ : Mk → Mni such that µ(eij) = fij<br />

for all i, j. By Lemma 2.6 there is then also a unital standard homomorphism<br />

Mk → Mni . But this is only possible if k|ni, cf. condition (2.9). <br />

Proof. Proof of Theorem 2.9 a) : By assumption A = lim(Mni<br />

−→ , ϕi) and B =<br />

lim<br />

−→ (Mmi , ψi) are the inductive limits of the two unital sequences (Mni , ϕi) and<br />

(Mmi , ψi), respectively. We will construct by induction two sequences k1 < k2 < . . .<br />

and l1 < l2 < . . . in N and unital ∗-homomorphisms µi : Mnk i → Mml i and<br />

λi : Mml i → Mnk i+1 such that<br />

ϕk2 ,k ϕk 1 3 ,k ϕk<br />

Mnk<br />

2 4 ,k<br />

Mnk<br />

3<br />

Mnk 1<br />

2<br />

3<br />

µ1<br />

<br />

Mml 1<br />

<br />

λ1 <br />

<br />

<br />

<br />

<br />

ψl 2 ,l 1<br />

µ2<br />

<br />

Mml 2<br />

<br />

λ2 <br />

<br />

<br />

<br />

<br />

ψl 3 ,l 2<br />

µ3<br />

<br />

Mml 3<br />

<br />

·<br />

· ·<br />

<br />

λ3 <br />

<br />

<br />

<br />

<br />

· · ·<br />

ψl 4 ,l 3<br />

commutes. So assume that we have constructed everything up to


54 1. FUNDAMENTALS<br />

ϕkr,kr−1 Mnk<br />

Mnkr<br />

r−1<br />

µr−2<br />

<br />

<br />

λr−1 <br />

<br />

<br />

<br />

<br />

µr−1<br />

Mml<br />

<br />

Mmlr r−1<br />

ψlr,lr−1 .<br />

Write mlr as a product of prime powers, say as<br />

mlr = p x1<br />

1 px2 2 · · ·pxs s .<br />

Then FB(pi) ≥ xi, i = 1, 2, . . ., s, by Corollary 2.14. But FA(pi) = FB(pi) by<br />

assumption, so FA(pi) ≥ xi for all i. There are therefore unital ∗-homomorphisms<br />

Mpi xi → A for all i. Another application of Corollary 2.14, now to A, shows that<br />

there are numbers ai ∈ N such that p xi<br />

i |nai for all i = 1, 2, . . ., s. Since nj|nj+1 for<br />

all j there is a number kr+1 > kr such that p xi<br />

i |nkr+1 for all i. But then<br />

mlr|nkr+1<br />

and hence there is a unital ∗-homomorphism λ : Mmlr → Mnk r+1 , for example the<br />

standard homomorphism determined by the 1 × 1 matrix ( nkr+1 ). It follows from<br />

mlr<br />

Lemma 2.6 that there is a unitary v ∈ Mnk such that Ad v ◦ λ ◦ µr−1 = ϕkr+1,kr.<br />

r+1<br />

We set λr = Ad v ◦ λ.<br />

To construct lr+1 and µr+1 we repeat the previous argument with the roles of<br />

A and B exchanged : Because FA = FB we see that there is a lr+1 > lr such that<br />

nkr+1|mlr+1 and hence such that there is a unital ∗-homomorphism µ : Mnk →<br />

r+1<br />

Mml r+1 . Again we conclude from Lemma 2.6 that there is a unitary w ∈ Mml r+1<br />

such that Ad w ◦ µ ◦ λr = ψlr,lr+1 and we set µr+1 = Ad w ◦ µ. Thus we get the<br />

following commutative diagram<br />

ϕkr+1 ,kr<br />

Mnkr<br />

<br />

Mnkr+1 µr<br />

<br />

<br />

λr <br />

<br />

<br />

<br />

<br />

µr+1<br />

Mmlr<br />

<br />

Mml .<br />

ψl<br />

r+1<br />

r+1 ,lr<br />

This completes the induction step.<br />

Having the infinite commutative diagram above we conclude from Proposition<br />

2.2 that A ≃ B.<br />

b) Let p1 < p2 < p3 < . . . be a numbering of P. For each n ∈ N we choose<br />

a n 1 ≤ a n 2 ≤ . . . in N such that limj→∞ a n j = F(pn). Set<br />

ri = p a1 i<br />

1 p a2 i<br />

2 · · ·p ai i<br />

i .<br />

Then ri|ri+1 and hence there is a unital ∗-homomorphism ϕi : Mri → Mri+1 . Let A<br />

be the inductive limit of the sequence<br />

Mr1<br />

ϕ1 <br />

Mr2<br />

ϕ2 <br />

Mr3<br />

ϕ3 <br />

. . .<br />

It is straightforward to check, by use of Corollary 2.14, that F = FA.


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 55<br />

It follows from Theorem 2.9 that there are uncountably many mutually nonisomorphic<br />

UHF-<strong>algebras</strong>. The invariant FA is sometimes refered to as a supernatural<br />

number because FA can be represented by a formal infinite product of prime powers:<br />

p a1<br />

1 p a2<br />

2 p a3<br />

3 · · ·<br />

where ai = FA(pi). In this terminology the UHF-algebra with Bratteli diagram<br />

is called the UHF-algebra of type 3 ∞ .<br />

·<br />

·<br />

·<br />

·<br />

.<br />

It is not always clear at first glance if a given C ∗ -algebra is an AF-algebra. It is<br />

therefore often useful to have a criterion which does not refer to inductive limits or<br />

increasing sequences of sub<strong>algebras</strong>. We will conclude this first introduction to AF<strong>algebras</strong><br />

by giving such a criterion. A C ∗ -algebra A is separable when it is separable<br />

as a metric space, i.e. if it contains a dense sequence. It is easy to see that a finitedimensional<br />

C ∗ -algebra is separable and that the inductive limit of a sequence of<br />

separable C ∗ -<strong>algebras</strong> is again separable. So an AF-algebra is certainly separable.<br />

This condition is not enough, however, as for example the case A = C[0, 1] shows.<br />

Lemma 2.15. Let n1, n2, . . .,nN ∈ N and ǫ > 0 be given. There is then a δ > 0<br />

with the following property : When {f d ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N} are<br />

matrix units in a <strong>C∗</strong>-algebra A and B is a <strong>C∗</strong>-subalgebra of A containing elements<br />

{xd ij } such that xdij − fd ij < δ for all i, j, d, then there are matrix units {edij : i, j =<br />

1, 2, . . ., nd, d = 1, 2, . . ., N} in B such that<br />

x d ij − edij < ǫ<br />

for all i, j, d and a unitary u ∈ Â such that<br />

and<br />

for all i, j, d.<br />

u − 1 < ǫ<br />

uf d ij u∗ = e d ij<br />

Proof. The following is only a sketch of the proof. In Exercise 2.30 below the<br />

reader is asked to provide the missing details.<br />

By substituting each xd ii by 1<br />

2 (xdii+x d ∗ d<br />

ii ), we may assume that each xii is selfadjoint.<br />

If δ is small enough, Lemma 2.10 gives us a projection E ∈ B close to <br />

Ex d ij E will be close to fd ij for all i, j, d. So we may assume that xd ij<br />

i,d fd ii. Then<br />

∈ EBE for all


56 1. FUNDAMENTALS<br />

i, j, d. If δ is small enough Lemma 2.10 gives us projections ed 11 ∈ EBE close to xd11 and hence also close to fd 11 . Then (E − ed11 )xd22 (E − ed11 ) is close to fd 22 and hence, if<br />

δ is small enough, Lemma 2.10 will give us a projection ed 22 ∈ (E − ed11 )A(E − ed 11 )<br />

close to (E − ed 11)xd 22(E − ed 11) and hence close to fd 22. Note that ed 11 and ed 22 are<br />

orthogonal. Then (E − ed 11 − ed 11)xd 33(E − ed 11 − ed 11) is close to fd 33 and we get in the<br />

same way a projection ed 33 ∈ B, orthogonal to both ed11 and ed22 , such that ed33 is close<br />

to fd 33 . By proceding in this way we get projections edii ∈ B such that edii ed′ kk = 0<br />

unless k = i, d = d ′ , and such that ed ii is close to fd ii. Furthermore, provided δ is<br />

small enough, Lemma 2.11 gives us partial isometries vd i ∈ B close to xd1i such that<br />

vd i vd∗ d<br />

i = eii, vd∗ d<br />

i vi = ed 11 for all i. Set ed ij = vd i vd∗ d<br />

j ∈ B and note that {eij} is a set<br />

of matrix units in B such that each ed ij is close to fd ij , say ed ij − fd ij < ǫ, for all<br />

i, j, d, if just δ is small enough. In particular, if only fd 11 − ed11 is small enough for<br />

all d, Lemma 2.11 (applied with x = ed 11fd 11 ) gives us partial isometries wd ∈ B such<br />

that wd − fd 11 is small and wdwd ∗ = ed 11, wd ∗wd = fd 11 for all d. Furthermore, since<br />

1 − <br />

i,d fd ii − (1 − <br />

i,d ed ii) is small, a similar argument provides us with a partial<br />

isometry v ∈ Â such that vv∗ = 1− <br />

i,d ed ii , v∗ v = 1− <br />

i,d fd ii<br />

and v−(1−<br />

i,d ed ii )<br />

is small. Set u = <br />

i,d ed i1 wdf d 1i + v. Then u is a unitary in  such that ufd ij u∗ = e d ij<br />

for all i, j, d. If only δ was small enough to begin with, we get in addition that<br />

u − 1 < ǫ. <br />

Theorem 2.16. A separable C ∗ -algebra A is an AF-algebra if and only if the<br />

following holds:<br />

For every finite set F ⊆ A and every ǫ > 0, there is a finite dimensional C ∗ -<br />

subalgebra B in A such that<br />

dist(x, B) < ǫ<br />

for all x ∈ F.<br />

Proof. It is clear from the definition of an AF-algebra that the condition is<br />

necessary: Since A = <br />

n An, where each An is finite dimensional, there is for each<br />

x ∈ F an integer nx ∈ N such that dist(x, Anx) < ǫ. Set n = maxx nx. Then<br />

dist(x, An) < ǫ for all x ∈ F.<br />

The problem is to prove the sufficiency of the condition. To this end we choose<br />

a dense sequence {yn} in A. Set xn = (max{yn, 1}) −1 yn, n ∈ N. Then {xn} is<br />

a dense sequence in the unit ball of A. We will construct a sequence F1 ⊆ F2 ⊆<br />

F3 ⊆ . . . of finite dimensional <strong>C∗</strong>-sub<strong>algebras</strong> of A such that dist(xj, Fn) < 1<br />

, j =<br />

n<br />

1, 2, . . ., n, n ∈ N. This will complete the proof because it will then be obvious that<br />

A = <br />

n Fn.<br />

F1 is found by using the assumption on A with F = {x1} and ǫ = 1. Assume<br />

that F1, F2, . . ., Fn have been constructed. Let {fd ij : i, j = 1, 2, . . ., nd, d =<br />

1, 2, . . ., N} be a set of matrix units which span Fn. Choose δ > 0 as in Lemma<br />

. By assumption there is a<br />

2.15, corresponding to {n1, n2, . . .,nN} and ǫ = 1<br />

5(n+1)<br />

finite dimensional <strong>C∗</strong>-subalgebra F of A such that dist(x, F) < min{δ, ǫ} for all<br />

x ∈ {x1, x2, . . .,xn+1} ∪ {fd ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N}. By Lemma 2.15<br />

there are matrix units {ed ij : i, j = 1, 2, . . ., nd, d = 1, 2, . . ., N} in F and a unitary<br />

1<br />

u ∈ Â such that u − 1 < ǫ = 5(n+1) such that ufd iju∗ = ed ij for all i, j, d. Set<br />

Fn+1 = u∗Fu. Then Fn+1 is a finite dimensional <strong>C∗</strong>-subalgebra of A such that<br />

Fn ⊆ Fn+1. Furthermore, for every j ∈ {1, 2, . . ., n + 1}, there is an element fj ∈ F


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 57<br />

such that fj − xj < ǫ. Then yj = u∗fju ∈ Fn+1 and yj = fj ≤ xj + ǫ ≤ 2.<br />

In addition yj − xj ≤ u∗fju − fj + ǫ < 5ǫ = 1 for all j = 1, 2, . . ., n + 1. This<br />

n+1<br />

completes the induction step and hence the proof. <br />

Corollary 2.17. An inductive limit of a sequence of AF-<strong>algebras</strong> is itself an<br />

AF-algebra.<br />

Proof. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

be a sequence of AF <strong>algebras</strong> and let A∞ be the inductive limit of the sequence and<br />

ϕn,∞ : An → A∞ the canonical maps. If {x n i : i ∈ N} is a dense sequence in An,<br />

then {ϕ∞,n(x n i ) : i, n ∈ N} is a countable dense set in A∞. Hence A∞ is separable.<br />

Let {x1, x2, . . .,xm} be a finite set in A∞ and let ǫ > 0. Since <br />

n ϕ∞,n(An) is<br />

dense in A∞ there is an N ∈ N and elements yi ∈ AN such that ϕ∞,N(yi) − xi <<br />

ǫ<br />

2 , i = 1, 2, . . .,m. Since AN is AF, there is a finite dimensional <strong>C∗</strong>-subalgebra F<br />

of AN containing elements zi ∈ F such that zi − yi < ǫ,<br />

i = 1, 2, . . ., m. Then<br />

2<br />

ϕ∞,N(F) is a finite dimensional <strong>C∗</strong>-subalgebra of A∞ and xi − ϕ∞,N(zi) < ǫ, i =<br />

1, 2, . . ., m. Hence A∞ satisfies the criterion of Theorem 2.16 and is consequently<br />

an AF algebra. <br />

2.3. Exercises.<br />

Exercise 2.18. Let (An, ϕn) be a sequence of C ∗ -<strong>algebras</strong>. Assume that there<br />

is a C ∗ -algebra D and a sequence of ∗-homomorphisms λn : An → D such that<br />

λn+1 ◦ ϕn = λn for all n and with the universal property of Lemma 2.1, i.e. such<br />

that whenever ψn : An → B, n ∈ N, is a sequence of ∗-homomorphisms into a<br />

common C ∗ -algebra B, satisfying that ψn+1 ◦ ϕn = ψn for all n, then there is one<br />

and only one ∗-homomorphism ψ : D → B such that ψ ◦ λn = ψn for all n ∈ N.<br />

Prove that there is a ∗-isomorphism Φ : D → A∞ such that Φ ◦ λn = ϕ∞,n for<br />

all n ∈ N.<br />

Exercise 2.19. Let (An, ϕn) and (Bn, ψn) be two sequences of C ∗ -<strong>algebras</strong>.<br />

Assume that there are sequences k1 < k2 < k3 < . . . and m1 < m2 < m3 < . . . in<br />

N and ∗-homomorphisms µi : Ami → Bki such that µi+1 ◦ ϕmi+1,mi = ψki+1,ki ◦ µi for<br />

all i ∈ N, i.e. such that the infinite diagram<br />

ϕm2 ,m1 ϕm3 ,m<br />

Am1<br />

2<br />

Am2<br />

Am3<br />

<br />

µ1<br />

Bk1 ψk 2 ,k 1<br />

<br />

µ2<br />

<br />

µ3<br />

ϕm 4 ,m 3<br />

. . .<br />

<br />

Bk2<br />

<br />

Bk3<br />

ψk3 ,k ψk 2 4 ,k3 commutes.<br />

Show that there is a unique ∗-homomorphism µ : A∞ → B∞ such that µ◦ϕ∞,mi =<br />

ψ∞,ki ◦ µi for all i ∈ N.<br />

Exercise 2.20. Let (An, ϕn) be a unital sequence of <strong>C∗</strong>-<strong>algebras</strong>. Show that<br />

A∞ = lim(An,<br />

ϕn) is unital with unit 1 = ϕ∞,1(1).<br />

−→<br />

Exercise 2.21. Set An = C[0, 1<br />

n ], n ∈ N, and define ϕn : C[0, 1<br />

n ] → C[0, 1<br />

n+1 ]<br />

by<br />

Determine lim<br />

−→ (An, ϕn).<br />

ϕn(f) = f| [0, 1<br />

n+1 ].<br />

<br />

. . .


58 1. FUNDAMENTALS<br />

Exercise 2.22. Let p, q be projections in the unital <strong>C∗</strong>-algebra A such that<br />

p − q < 1. Show that there is a unitary u ∈ A such that upu∗ = q. (Hint : Set<br />

u = v(v∗v) −1<br />

2 where v = 1 − p − q + 2pq.)<br />

Exercise 2.23. Let X = {0} ∪ { 1 : n ∈ N, n ≥ 1}, considered as a compact<br />

n<br />

subset of [0, 1]. Show that C(X) is a unital AF-algebra with Bratteli diagram<br />

etc.<br />

· <br />

<br />

<br />

· · <br />

<br />

<br />

· · · <br />

<br />

<br />

· · · · <br />

<br />

Exercise 2.24. Let F1 = [0, 1], F2 = [0, 1<br />

.<br />

.<br />

.<br />

.<br />

3 ]∪[2<br />

.<br />

3 , 1], F3 = [0, 1<br />

9 ]∪[2 1 , 9 3 ]∪[2 7 , 3 9 ]∪[8 , 1], 9<br />

Then K = ∞<br />

n=1 Fn is the (middle third) Cantor set. Show that C(K) is a unital<br />

AF-algebra with Bratteli diagram<br />

· <br />

<br />

<br />

· ·<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

. . . .<br />

Exercise 2.25. Let A be the unital AF-algebra with Bratteli diagram


2. INDUCTIVE LIMITS OF C ∗ -ALGEBRAS 59<br />

·<br />

<br />

<br />

<br />

· · ·<br />

<br />

<br />

·<br />

<br />

<br />

<br />

· · ·<br />

<br />

<br />

·<br />

<br />

<br />

<br />

· · ·<br />

<br />

<br />

·<br />

<br />

<br />

. . .<br />

Show that A is isomorphic to the UHF-algebra of type 3 ∞ .<br />

Exercise 2.26. Let A be the AF-algebra with Bratteli diagram<br />

1<br />

<br />

2<br />

<br />

3<br />

<br />

4<br />

<br />

5<br />

<br />

.<br />

Show that A ≃ K, the compact operators on an infinite dimensional separable<br />

Hilbert space


60 1. FUNDAMENTALS<br />

Exercise 2.27. Let A be the unital AF-algebra with Bratteli diagram<br />

·<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· ·<br />

<br />

<br />

<br />

<br />

<br />

<br />

. . .<br />

Show that A is isomorphic to the UHF-algebra of type 2 ∞ .<br />

Exercise 2.28. Let A be the unital AF-algebra with diagram<br />

·<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· <br />

·<br />

<br />

<br />

<br />

<br />

<br />

· ·<br />

<br />

<br />

<br />

<br />

<br />

<br />

. . .<br />

Show that A is not isomorphic to a UHF-algebra, for example by showing that A<br />

does not contain a full matrix algebra Mn for any n ≥ 2 as a unital C ∗ -subalgebra.<br />

Exercise 2.29. Show that the C ∗ -algebra C[0, 1] is separable but not an AFalgebra.<br />

Exercise 2.30. Fill in the details missing in the proof of Lemma 2.15.<br />

Exercise 2.31. Let A be a unital separable C ∗ -algebra with the following property<br />

: For any finite set F ⊆ A and for any ǫ > 0 there is a C ∗ -subalgebra B ⊆ A<br />

such that B is ∗-isomorphic to Mn for some n ∈ N and dist(x, B) < ǫ for all x ∈ F.<br />

Show that A is a UHF-algebra.


y<br />

3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 61<br />

Exercise 2.32. Define a unital ∗-homomorphism ϕn : C([0, 1], M2 n) → C([0, 1], M 2 n+1)<br />

ϕn(f)(t) =<br />

<br />

t f( ) 0<br />

2<br />

0 f( t+1<br />

2 )<br />

Show that A∞ = lim(C([0,<br />

1], M2 −→ n), ϕn) is an AF-algebra. (Hint : Let a ∈ C([0, 1], M2n+1). Show that for every ǫ > 0, there is a k ≥ n such that ϕk,n(a) ∈ C([0, 1], M2k+1) is<br />

within the distance ǫ from an element of M2k+1 ⊆ C([0, 1], M2k+1). Then use Theorem<br />

2.16.) Try to show that A∞ is a UHF-algebra of type 2∞ .<br />

<br />

, t ∈ [0, 1] .<br />

3. K0 and the classification of AF-<strong>algebras</strong><br />

3.1. Definition of K0. Let A be a C ∗ -algebra.<br />

Definition 3.1. Two projections p, q ∈ A are equivalent when there is a partial<br />

isometry v ∈ A such that vv ∗ = p, v ∗ v = q. We write p ≈v q in this case, or just<br />

p ≈ q when it is not necessary to specify the partial isometry.<br />

Lemma 3.2. 1) ≈ is an equivalence relation on the set of projections in A.<br />

2) When p, q, p1, q1 are projections in A such that pp1 = qq1 = 0 and p ≈ q, p1 ≈ q1,<br />

then p + p1 ≈ q + q1.<br />

3) When p, q ∈ A are projections such that p ≈ q and 1 − p ≈ 1 − q (in Â),<br />

then there is a unitary u ∈ Â such that upu∗ = q.<br />

4) When p, q ∈ A are projections in A and u ∈ Â is a unitary such that<br />

upu∗ = q, then p ≈ q and 1 − p ≈ 1 − q (in Â).<br />

Proof. 1) : Since p ≈p p, ≈ is a reflexive relation. Since p ≈v q ⇒ q ≈v∗ p, the<br />

relation is also symmetric. Finally, if p ≈v q, q ≈w r, then p ≈vw r. Indeed,<br />

while<br />

Thus<br />

vw(vw) ∗ = vww ∗ v ∗ = vqv ∗ = vv ∗ vv ∗ = p 2 = p,<br />

(vw) ∗ vw = w ∗ v ∗ vw = w ∗ qw = w ∗ ww ∗ w = r 2 = r.<br />

2): If p ≈v q and p1 ≈w q1, note first that vw ∗ = 0 because<br />

vw ∗ (vw ∗ ) ∗ = vw ∗ wv ∗ = vq1v ∗ = vv ∗ vq1v ∗ = vqq1v ∗ = 0.<br />

(v + w)(v + w) ∗ = vv ∗ + vw ∗ + wv ∗ + ww ∗ = vv ∗ + ww ∗ = p + p1.<br />

Similarly, (v + w) ∗ (v + w) = q + q1 because v ∗ w = 0.<br />

3) : If p ≈v q and 1 − p ≈w 1 − q, then 1 ≈v+w 1 by 2); i.e. v + w is a unitary<br />

and<br />

(v + w)q(v + w) ∗ = vqv ∗ + vqw ∗ + wqv ∗ + wqw ∗ =<br />

vv ∗ vv ∗ + vq(1 − q)w ∗ + w(1 − q)qv ∗ + w(1 − q)qw ∗ = vv ∗ vv ∗ = p 2 = p.<br />

4) It is straightforward to check that q ≈up p. <br />

Let Pn(A) denote the set of projections in Mn(A) and let P(A) denote the disjoint<br />

union<br />

∞<br />

P(A) = Pn(A).<br />

n=1


62 1. FUNDAMENTALS<br />

When e ∈ Pn(A), f ∈ Pk(A) we define e ⊕ f to be the projection<br />

<br />

e 0<br />

0 f<br />

in Pn+k(A). Also, we extend the equivalence relation ≈ to P(A) by setting e≈f<br />

when <br />

e 0 f 0<br />

≈<br />

0 0 0 0<br />

in Mm(A) for some m ≥ k, n. We leave the reader to check that ≈ is an equivalence<br />

relation also on P(A) and that it is indeed an extension of the equivalence relation<br />

≈ in P1(A); more precisely, that two projections p, q ∈ A are equivalent in A if and<br />

only if they are equivalent in P(A).<br />

Lemma 3.3. Let<br />

⎛<br />

⎞<br />

p11 p12 p13 . . . p1n<br />

⎜p21<br />

p22 p23 . . . p2n⎟<br />

⎜<br />

⎟<br />

p = ⎜p31<br />

p32 p33 . . . p3n⎟<br />

⎜<br />

⎝ .<br />

⎟ ∈ Pn(A).<br />

. . . .. . ⎠<br />

pn1 pn2 pn3 . . . pnn<br />

Then<br />

⎛<br />

p11<br />

⎜p21<br />

⎜<br />

⎜p31<br />

⎜ .<br />

⎝pn1<br />

p12<br />

p22<br />

p32<br />

.<br />

pn2<br />

p13<br />

p23<br />

p33<br />

.<br />

pn3<br />

. . . p1n<br />

. . . p2n<br />

. . . p3n<br />

. .. .<br />

. . . pnn<br />

⎞ ⎛<br />

0 0 0<br />

0⎟<br />

⎜0<br />

p11 ⎟ ⎜<br />

0⎟<br />

⎜0<br />

p21<br />

⎟<br />

.<br />

⎟ ≈ ⎜<br />

⎜0<br />

p31<br />

⎟ ⎜<br />

0⎠<br />

⎝.<br />

.<br />

0<br />

p12<br />

p22<br />

p32<br />

.<br />

0<br />

p13<br />

p23<br />

p33<br />

.<br />

⎞<br />

. . . 0<br />

. . . p1n ⎟<br />

. . . p2n ⎟<br />

. . . p3n⎟<br />

.<br />

⎟<br />

.. . ⎠<br />

0 0 0 . . . 0 0 0 pn1 pn2 pn3 . . . pnn<br />

in Mn+1(A).<br />

Proof. Set<br />

⎛<br />

⎞<br />

0 0 0 . . . 0 1<br />

⎜1<br />

0 0 . . . 0 0⎟<br />

⎜<br />

⎟<br />

⎜0<br />

1 0 . . . 0 0⎟<br />

W = ⎜<br />

⎟<br />

⎜0<br />

0 1 . . . 0 0⎟<br />

∈ Mn+1(<br />

⎜<br />

⎝ ..<br />

⎟<br />

. . . . . . ⎠<br />

0 0 0 . . . 1 0<br />

Â).<br />

Then W is a unitary in Mn+1( Â) and a direct calculation shows that<br />

⎛<br />

⎞<br />

p11 p12 p13 . . . p1n 0<br />

⎜p21<br />

p22 p23 . . . p2n 0⎟<br />

⎜<br />

⎟<br />

⎜p31<br />

p32 p33 . . . p3n 0⎟<br />

W ⎜ .<br />

⎟<br />

⎜ . . . .. . .<br />

⎟W<br />

⎟<br />

⎝pn1<br />

pn2 pn3 . . . pnn 0⎠<br />

0 0 0 . . . 0 0<br />

∗ ⎛<br />

⎞<br />

0 0 0 0 . . . 0<br />

⎜0<br />

p11 p12 p13 . . . p1n⎟<br />

⎜<br />

⎟<br />

⎜0<br />

p21 p22 p23 . . . p2n⎟<br />

= ⎜<br />

⎟<br />

⎜0<br />

p31 p32 p33 . . . p3n⎟<br />

.<br />

⎜<br />

⎝<br />

.<br />

⎟<br />

. . . . .. . ⎠<br />

0 pn1 pn2 pn3 . . . pnn<br />

Lemma 3.4. 1) If e1, e2, f1, f2 ∈ P(A) with e1 ≈ e2 and f1 ≈ f2, then<br />

e1 ⊕ f1 ≈ e2 ⊕ f2.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 63<br />

2) e ⊕ f ≈ f ⊕ e for all e, f ∈ P(A).<br />

3) e ⊕ (f ⊕ g) ≈ (e ⊕ f) ⊕ g for all e, f, g ∈ P(A).<br />

Proof. 1) By assumption there is a k such that<br />

<br />

e1 0 e2 0<br />

≈<br />

0 0 0 0<br />

within Mk(A) and<br />

<br />

f1<br />

0<br />

<br />

0 f2<br />

≈<br />

0 0<br />

<br />

0<br />

,<br />

0<br />

also within Mk(A). It follows from Lemma 3.2 2) that<br />

⎛ ⎞<br />

e1 0 0 0<br />

⎜ 0 0 0 0 ⎟<br />

⎝ 0 0 f1 0⎠<br />

0 0 0 0<br />

≈<br />

⎛<br />

e2 0 0<br />

⎜ 0 0 0<br />

⎝ 0 0 f2<br />

⎞<br />

0<br />

0 ⎟<br />

0⎠<br />

0 0 0 0<br />

within M2k(A). But repeated use of Lemma 3.3 shows that<br />

⎛ ⎞<br />

e1 0 0 0<br />

⎜ 0 0 0 0 ⎟<br />

⎝ 0 0 f1 0⎠<br />

0 0 0 0<br />

≈<br />

⎛ ⎞<br />

e1 0 0 0<br />

⎜ 0 f1 0 0 ⎟<br />

⎝ 0 0 0 0⎠<br />

0 0 0 0<br />

≈ e1 ⊕ f1<br />

and ⎛<br />

e2<br />

⎜ 0<br />

⎝ 0<br />

0 0<br />

0 0<br />

0 f2<br />

⎞<br />

0<br />

0 ⎟<br />

0⎠<br />

0 0 0 0<br />

≈<br />

⎛<br />

e2<br />

⎜ 0<br />

⎝ 0<br />

0<br />

f2<br />

0<br />

⎞<br />

0 0<br />

0 0 ⎟<br />

0 0⎠<br />

0 0 0 0<br />

≈ e2 ⊕ f2<br />

To prove 2) observe first that we may assume that e, f ∈ Pk(A) for the same k by<br />

1). Set<br />

<br />

0 1<br />

W = ∈ M2k(<br />

1 0<br />

Â).<br />

W is a unitary such that<br />

<br />

e 0<br />

W W<br />

0 f<br />

∗ <br />

f 0<br />

= .<br />

0 e<br />

3) is a triviality since in fact e ⊕ (f ⊕ g) = (e ⊕ f) ⊕ g.<br />

It follows from Lemma 3.4 that set of equivalence classes in P(A) form an Abelian<br />

semigroup when we set<br />

[e] + [f] = [e ⊕ f], e, f ∈ P(A).<br />

This semigroup, P(A)/≈, will be denoted by V (A). We shall now apply a standard<br />

algebraic trick - the Grothendieck construction - to produce a group, K00(A), from<br />

V (A). First we describe the Grothendieck construction.<br />

Let S be an Abelian semi-group, i.e. S is a set endowed with a composition +<br />

such that s + t = t + s and (s + t) + u = s + (t + u) for all s, t, u ∈ S. Define an<br />

equivalence relation ∼ on S × S by<br />

(x, u) ∼ (y, v) ⇔ there is an element r ∈ S such that x + v + r = y + u + r.


64 1. FUNDAMENTALS<br />

We leave the reader to check that this is indeed an equivalence relation on S ×S.<br />

The equivalence class containing (x, y) ∈ S × S will be denoted by [x, y]. The set of<br />

equivalence classes, S × S/ ∼, will be denoted by G(S). Define a composition + in<br />

G(S) by<br />

[x, y] + [u, v] = [x + u, y + v].<br />

Then G(S) is an Abelian group with [x, x] (for any x ∈ S) as neutral element and<br />

with −[x, y] = [y, x]. (CHECK !) There is a canonical semi-group homomorphism<br />

ϕ : S → G(S) given by<br />

ϕ(x) = [x + r, r]<br />

for some fixed r ∈ S. The crucial functorial property of this construction is described<br />

in the following<br />

Proposition 3.5. If H is an Abelian group and µ : S → H a semi-group<br />

homomorphism, then there is a unique group homomorphism ˆµ : G(S) → H such<br />

that ˆµ ◦ ϕ = µ.<br />

Proof. Note first that [x, y] = [x + r, r] − [y + r, r] for all x, y ∈ S. Thus<br />

G(S) = ϕ(S) − ϕ(S) and we see that any homomorphism G(S) → H is determined<br />

by its restriction to ϕ(S). This gives the uniqueness of ˆµ. To construct ˆµ, set<br />

ˆµ([x, y]) = µ(x) − µ(y), x, y ∈ S. It is straightforward to check that ˆµ is a welldefined<br />

group homomorphism. Since ˆµ ◦ ϕ(x) = ˆµ([x + r, r]) = µ(x + r) − µ(r) =<br />

µ(x), x ∈ S, the proof is complete. <br />

Definition 3.6. We define K00(A) to be the group G(V (A)), i.e. the group<br />

obtained by applying the Grothendieck construction to the semi-group V (A) of<br />

equivalence classes of projections in P(A).<br />

For every e ∈ P(A) we denote by [e] the element of K00(A) which is the image<br />

of [e] ∈ V (A) under the canonical homomorphism ϕ : V (A) → G(V (A)) = K00(A).<br />

We summarize the most basic aspects of the construction of K00(A) as follows.<br />

Lemma 3.7. 1) Every element of K00(A) has the form [e] − [f] for some<br />

e, f ∈ P(A).<br />

2) If e1, e2, f1, f2 ∈ P(A), then [e1] − [f1] = [e2] − [f2] in K00(A) if and only if<br />

there is a q ∈ P(A) such that<br />

e1 ⊕ f2 ⊕ q ≈ e2 ⊕ f1 ⊕ q.<br />

3) ([e1] − [f1]) + ([e2] − [f2]) = [e1 ⊕ e2] − [f1 ⊕ f2] for all e1, e2, f1, f2 ∈ P(A).<br />

4) The zero element of K00(A) is [0].<br />

We now develop the basic functorial properties of the K00-construction. So let B<br />

be another <strong>C∗</strong>-algebra and ψ : A → B a ∗-homomorphism. ψ extends in a canonical<br />

way to a ∗-homomorphism Mn(A) → Mn(B) which we also denote by ψ, namely<br />

⎛<br />

⎞<br />

a11 a12 . . . a1n<br />

⎜a21<br />

a22 . . . a2n⎟<br />

ψ ⎜<br />

⎝ .<br />

⎟<br />

. . .. .<br />

⎠<br />

an1 an2 . . . ann<br />

=<br />

⎛<br />

⎞<br />

ψ(a11) ψ(a12) . . . ψ(a1n)<br />

⎜ψ(a21)<br />

ψ(a22) . . . ψ(a2n) ⎟<br />

⎜<br />

⎝<br />

.<br />

⎟<br />

. . .. .<br />

⎠<br />

ψ(an1) ψ(an2) . . . ψ(ann)<br />

.<br />

In this way ψ induces a map ψ : P(A) → P(B) which in turn defines a map<br />

V (A) → V (B) given by<br />

[e] ↦→ [ψ(e)], e ∈ P(A).


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 65<br />

It is clear that this is a semi-group homomorphism, so if we compose it with the<br />

canonical map V (B) → K00(B) we can use Proposition 3.5 to conclude that there<br />

is a group homomorphism<br />

such that<br />

ψ∗ : K00(A) → K00(B)<br />

ψ∗([e] − [f]) = [ψ(e)] − [ψ(f)], e, f ∈ P(A).<br />

In this way the construction of the K00-group becomes a functor from C ∗ -<strong>algebras</strong><br />

to Abelian groups. In particular,<br />

Lemma 3.8. Let ψ : A → B and ϕ : B → C be ∗-homomorphisms between<br />

C ∗ -<strong>algebras</strong> then<br />

(ϕ ◦ ψ)∗ = ϕ∗ ◦ ψ∗.<br />

Although this lemma is trivial, it is the core of the usefulness of the whole<br />

construction because it implies immediately that two isomorphic C ∗ -<strong>algebras</strong> must<br />

have isomorphic K00-groups. In the rest of this section we develop the K00-theory<br />

for the purpose of using it for the functor which interests us more, namely K0.<br />

Lemma 3.9. Let ϕ, ψ : A → B be ∗- homomorphisms between C ∗ -<strong>algebras</strong> such<br />

that<br />

ϕ(A)ψ(A) = {0}.<br />

Then ϕ + ψ : A → B is a ∗-homomorphism and (ϕ + ψ)∗ = ϕ∗ + ψ∗ : K00(A) →<br />

K00(B).<br />

Proof. It is straightforward to check that ϕ + ψ is a ∗-homomorphism. To<br />

obtain the formula for (ϕ + ψ)∗ note first that if e, f are orthogonal projections in<br />

Mn(B) then<br />

in M2n(B) with<br />

<br />

e + f 0<br />

0 0<br />

v =<br />

≈v<br />

<br />

e f<br />

0 0<br />

<br />

e 0<br />

0 f<br />

So if p is a projection in Mn(A) then ϕ(p) and ψ(p) are orthogonal projections in<br />

Mn(B) and hence<br />

<br />

ϕ(e) 0 ϕ(e) + ψ(e) 0<br />

[ϕ(e)] + [ψ(e)] = [ ] = [<br />

] = (ϕ + ψ)∗([e])<br />

0 ψ(e) 0 0<br />

Thus (ϕ∗ + ψ∗)[e] = (ϕ + ψ)∗([e]) and the desired conclusion follows because of<br />

Lemma 3.7 1).<br />

<br />

Lemma 3.10. Let A and B be C ∗ -<strong>algebras</strong>. Then<br />

K00(A ⊕ B) ≃ K00(A) ⊕ K00(B).<br />

More precisely, if π A : A ⊕ B → A and π B : A ⊕ B → B denote the two projections,<br />

then the map K00(A ⊕ B) ∋ x ↦→ (π A ∗ (x), πB ∗ (x)) ∈ K00(A) ⊕ K00(B) is an<br />

isomorphism.


66 1. FUNDAMENTALS<br />

Proof. Let iA : A → A ⊕ B and iB : B → A ⊕ B be the maps iA(a) = (a, 0)<br />

and iB(b) = (0, b), respectively. Then π A ◦ iA = idA and π B ◦ iB = idB. If γ denotes<br />

the map from the statement of the lemma and µ : K00(A) ⊕ K00(B) → K00(A ⊕ B)<br />

the map µ(x, y) = iA∗(x) + iB∗(y), then<br />

And<br />

µ ◦ γ(x) = (iA ◦ π A )∗(x) + (iB ◦ π B )∗(x)<br />

= (iA ◦ π A + iB ◦ π B )∗(x) (by Lemma 3.9)<br />

= idA⊕B∗ (x) = idK00(A⊕B)(x) = x ( because iA ◦ π A + iB ◦ π B = idA⊕B).<br />

γ ◦ µ(x, y) = γ(iA∗(x) + iB∗(y)) = (π A ◦ iA)∗(x), (π B ◦ iB)∗(y) = (x, y).<br />

For the purpose of classifying C ∗ -<strong>algebras</strong> it is important to keep track of the<br />

semigroup V (A) which gave us the group K00(A). Thus we set<br />

K00(A) + = {[e] : e ∈ P(A)}. (3.1)<br />

In other words, K00(A) + is the image of V (A) in K00(A) under the canonical map<br />

V (A) → K00(A) coming with the Grothendieck construction. In particular, K00(A) +<br />

is a semi-group in K00(A) and hence it gives rise to a notion of positivity within<br />

K00(A), namely we write x ≤ y when y − x ∈ K00(A) + .<br />

To study the notion of positivity we first consider the general notion of preordered<br />

Abelian groups.<br />

Definition 3.11. A pre-ordered Abelian group G is an Abelian group endowed<br />

with a reflexive, transitive and translation invariant relation ≤. In other words,<br />

1) g ≤ g,<br />

2) g ≤ h, h ≤ k ⇒ g ≤ k,<br />

3) g ≤ h ⇒ g + k ≤ h + k.<br />

for all g, h, k ∈ G.<br />

In a pre-ordered Abelian group G the set G + = {x ∈ G : x ≥ 0} is a semigroup,<br />

called the positive cone in G. Note that G + determines the relation ≤ because<br />

x ≤ y ⇔ y − x ∈ G + . Conversely, a subsemigroup G + of G with 0 ∈ G + defines a<br />

relation x ≤ y on G by x ≤ y ⇔ y − x ∈ G + which makes G a pre-ordered Abelian<br />

group.<br />

A pre-ordered Abelian group is called partially ordered when G + ∩(−G + ) = {0},<br />

i.e. if x ≤ y, y ≤ x ⇒ x = y.<br />

It is now a triviality to check that<br />

Lemma 3.12. K00(A) is a pre-ordered Abelian group with K00(A) + = {[e] : e ∈<br />

P(A)} as positive cone.<br />

A homomorphism ϕ : G → H between pre-ordered Abelian groups is positive<br />

when ϕ(G + ) ⊆ H + . Note that the group homomorphism ϕ∗ : K00(A) → K00(B)<br />

induced by a ∗-homomorphism ϕ : A → B is positive.<br />

We will now construct the inductive limits in the category of pre-ordered Abelian<br />

groups. Let<br />

G1<br />

p1 <br />

G2<br />

p2 <br />

G3<br />

p3 <br />

. . .


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 67<br />

be a sequence of pre-ordered Abelian groups with positive connecting homomorphisms.<br />

Two elements g ∈ Gn, h ∈ Gm are equivalent, written g ∼ h, in the disjoint<br />

union<br />

∞<br />

i=1<br />

when pk,n(g) = pk,m(h) for some k ≥ n, m. The equivalence classes of this equivalence<br />

relation form an Abelian group G with composition<br />

[g] + [h] = [pk,n(g) + pk,m(h)], k ≥ n, m, g ∈ Gn, h ∈ Gm.<br />

The homomorphisms qn : Gn → G given by qn(g) = [g] will be called the canonical<br />

homomorphisms . We set<br />

G + ∞<br />

= qn(G + n)<br />

n=1<br />

which is a semigroup in G and we define the relation ≤ on G by x ≤ y ⇔ y−x ∈ G + .<br />

Proposition 3.13. (a) G is a pre-ordered Abelian group and, if each Gn is<br />

partially ordered, then so is G.<br />

(b) G is the inductive limit of the sequence (Gn, pn) in the category of pre-ordered<br />

Abelian groups, i.e. if H is a pre-ordered Abelian group and ri : Gi → H, i ∈ N,<br />

are positive group homomorphisms such that rn+1 ◦ pn = rn for all n, then there is<br />

a unique positive group homomorphism r : G → H such that r ◦ qn = rn.<br />

Gi<br />

Proof. (a) It is straightforward to check that G is a pre-ordered Abelian group.<br />

We show that G is partially ordered if each Gn is. So assume that x ∈ G + ∩(−G + ).<br />

It follows that there is a k ∈ N and elements g, h ∈ G +<br />

k<br />

x = qk(g), −x = qk(h).<br />

such that<br />

But then for some n ≥ k, −pn,k(h) = pn,k(g) ∈ G + n ∩ (−G+ n ) = {0} because Gn is<br />

partially ordered. Hence x = qn(pn,k(g)) = 0.<br />

(b) Since rn+1 ◦ pn = rn, we can define a homomorphism r : G → H by<br />

r([g]) = rn(g), g ∈ Gn, n ∈ N. Since r(qn(G + n)) = rn(G + n) ⊆ H + , r is a positive<br />

homomorphism. Since G = ∞<br />

n=1 qn(Gn), it is clear that r is uniquely determined<br />

by the condition that r ◦ qn = rn for all n. <br />

Note that the preceding construction gives the inductive limit of a sequence of<br />

Abelian groups (not necessarily pre-ordered) because we can always consider an<br />

Abelian group as being pre-ordered by the positive cone {0}.<br />

To describe the relation between inductive limits of C ∗ -<strong>algebras</strong> and inductive<br />

limits of pre-ordered Abelian groups we first need some lemmas. The first is an<br />

immediate corollary of Lemma 2.11.<br />

Lemma 3.14. Let p, q be projections in the <strong>C∗</strong>-algebra A. Assume that p−q ≤<br />

δ < 1<br />

3 . It follows that there is a partial isometry v ∈ A such that vv∗ = p, v∗v = q<br />

and v − pq ≤ 4 √ δ.<br />

Proof. Set x = pq and note that xx ∗ − p = pqp − p = p(q − p)p ≤<br />

q − p, x ∗ x − q = q(p − q)q ≤ q − p. Apply Lemma 2.11.


68 1. FUNDAMENTALS<br />

Lemma 3.15. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of C ∗ -<strong>algebras</strong> and A = lim<br />

−→ (An, ϕn) the inductive limit C ∗ -algebra.<br />

Let p be a projection in Md(A) for some d ∈ N. For any ǫ > 0 there is an m ∈ N<br />

and a projection q ∈ Md(Am) such that<br />

ϕ∞,m(q) − p < ǫ .<br />

Proof. For any δ > 0 we can find i ∈ N and y ∈ Md(Ai) such that<br />

ϕ∞,i(y) − p < δ.<br />

By using 1<br />

2 (y + y∗ ) instead of y we may assume that y = y∗ . If δ ≤ 1 we find<br />

that ϕ∞,i(y) ≤ 2 and ϕ∞,i(y2 − y) < 4δ. By using (1.25) this implies that<br />

ϕ∞,i((y2 − y)kl) < 4δ for all k, l ∈ {1, 2, . . ., d}. There is therefore an m > i such<br />

that ϕm,i((y2 − y)kl) < 4δ for all k, l. But then (1.25) implies that ϕm,i(y2 −<br />

y) < 4d 3<br />

2δ. Thus, if 4d 3<br />

2δ < 1<br />

4 , we have a projection q ∈ Md(Am) such that<br />

q − ϕm,i(y) < 8d 3<br />

2δ by Lemma 2.10. It follows that<br />

ϕ∞,m(q) − p < (8d 3<br />

2 + 1)δ.<br />

We have completed the proof if (8d 3<br />

2 + 1)δ < ǫ. <br />

Theorem 3.16. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of <strong>C∗</strong>-<strong>algebras</strong> and A = lim(An,<br />

ϕn) the inductive limit C<br />

−→ ∗-algebra. Let<br />

K00(A1) ϕ1∗ <br />

K00(A2) ϕ2∗ <br />

K00(A3) ϕ3∗ <br />

. . .<br />

be the sequence of pre-ordered Abelian groups obtained by applying the K00-functor<br />

and let G = lim(K00(An),<br />

ϕn∗) be the inductive limit in the category of pre-ordered<br />

−→<br />

Abelian groups and qn : K00(An) → G the canonical homomorphisms.<br />

There is then a unique positive homomorphism q : G → K00(A) such that q◦qn =<br />

ϕ∞,n∗ for all n ∈ N and q is an isomorphism in the category of pre-ordered Abelian<br />

groups (i.e. q is a group isomorphism such that both q and q−1 are positive.)<br />

Proof. Note that ϕ∞,n+1 ∗ ◦ ϕn∗ = ϕ∞,n ∗ for all n, so the existence of a unique<br />

positive homomorphism q : G → K00(A) with the property that q ◦ qn = ϕ∞,n ∗ for<br />

all n follows from Proposition 3.13 (b).<br />

Let x ∈ K00(A) + so that x = [e] for some projection e ∈ Mn(A). By Lemma<br />

3.14 and Lemma 3.15 there is an m ∈ N and a projection p ∈ Mn(Am) such that<br />

e ≈ ϕ∞,m(p). Hence<br />

x = [e] = [ϕ∞,m(p)] = ϕ∞,m ∗ ([p]) = q(qm([p])),<br />

proving that q(G + ) = K00(A) + . Since K00(A) = K00(A) + − K00(A) + by Lemma 3.7<br />

1), this shows also that q is surjective. To show that q is a positive isomorphism<br />

with positive inverse it is now sufficient to prove that q is injective.<br />

Consider x ∈ ker q. Since G = <br />

i qi(K00(Ai)), there are n, i ∈ N and projections<br />

p1, q1 ∈ Mn(Ai) such that x = qi([p1] − [q1]). Since<br />

0 = q(x) = ϕ∞,i ∗ ([p1] − [q1]) = [ϕ∞,i(p1)] − [ϕ∞,i(q1)]


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 69<br />

we know from Lemma 3.7 2) that there is a projection f ∈ P(A) such that ϕ∞,i(p1)⊕<br />

f ≈ ϕ∞,i(q1)⊕f. From the argument above we know that there is a j ≥ i, an m ∈ N<br />

and a projection h ∈ Mm(Aj) such that f ≈ ϕ∞,j(h) in Mm(A). Set k = n + m and<br />

p = ϕj,i(p1) ⊕ h, q = ϕj,i(q1) ⊕ h. Both p and q are projections in Mk(Aj) and we<br />

know that<br />

ϕ∞,j(p) ≈ ϕ∞,i(p1) ⊕ f ≈ ϕ∞,i(q1) ⊕ f ≈ ϕ∞,j(q)<br />

in Mk(A). There is therefore a partial isometry w ∈ Mk(A) such that<br />

ww ∗ = ϕ∞,j(p), w ∗ w = ϕ∞,j(q).<br />

Let δ ∈]0, 1[ and find a ∈ Mk(Al) for some l ≥ j such that ϕ∞,l(a) − w < δ.<br />

Then<br />

ϕ∞,l(aa ∗ − ϕl,j(p)) < 4δ, ϕ∞,l(a ∗ a − ϕl,j(q)) < 4δ.<br />

By using (1.25) as in the proof of Lemma 3.15 we conclude that there is a r ≥ l such<br />

that<br />

ϕr,l(aa ∗ − ϕl,j(p)) < 4k 3<br />

2δ, ϕr,l(a ∗ a − ϕl,j(q)) < 4k 3<br />

2δ .<br />

In particular, ϕr,l(a)2 ≤ ϕl,j(p))+4k 3<br />

2δ ≤ 1+4k 3<br />

2δ. If just δ > 0 is small enough<br />

we see that x = (1 + 4k 3<br />

2δ) −1 2ϕr,l(a) is an element of Mk(Ar) such that x ≤ 1 and<br />

xx ∗ − ϕr,j(p) < 1<br />

3 , x∗x − ϕr,j(q) < 1<br />

3 .<br />

It follows from Lemma 2.11 that ϕr,j(p) ≈ ϕr,j(q) in Mk(Ar), and hence that<br />

[ϕr,j(p)] = [ϕr,j(q)] in K00(Ar). We conclude that<br />

x = qi([p1] − [q1]) = qj([ϕj,i(p1)] − [ϕj,i(q1)]) = qj([ϕj,i(p1)] + [h] − ([ϕj,i(q1)] + [h])) =<br />

qj([p] − [q]) = qr([ϕr,j(p)] − [ϕr,j(q)]) = qr(0) = 0.<br />

Hence q is injective and the proof is complete. <br />

An alternative, shorter and maybe more instructive way to formulate the result<br />

in Theorem 3.16 is that<br />

K00(lim An) = lim K00(An)<br />

−→ −→<br />

as pre-ordered Abelian groups.<br />

To introduce the genuine K0-group we need to adjoin units to a <strong>C∗</strong>-algebra in<br />

a way which allows us to deal with unital and non-unital <strong>algebras</strong> in the same way.<br />

The method used to add a unit which we used in Section 1 is not sufficiently wellbehaved<br />

from a functorial point of view. The trouble is that a ∗-homomorphism<br />

ϕ : A → B may not extend to a unital ∗-homomorphism  → ˆ B. To repair this<br />

defect it is necessary to add a unit in a way which changes the algebra even when it<br />

is unital to begin with. We summarize the construction in the following<br />

Lemma 3.17. Let A be a C ∗ -algebra and endow the vector space direct sum A⊕C<br />

with the involution<br />

(a, λ) ∗ = (a ∗ , λ)<br />

and the product<br />

(a, λ)(b, µ) = (ab + λb + µa, λµ).<br />

The resulting ∗-algebra is a C ∗ -algebra A + with unit (0, 1) and there is a unital ∗homomorphism<br />

χA : A + → C given by χA(a, λ) = λ such that<br />

ker χA = A ⊆ A + .<br />

Furthermore,


70 1. FUNDAMENTALS<br />

1) When A is unital, then (a, λ) → (a+λ1, λ) defines an isomorphism between<br />

A + and the C ∗ - algebra direct sum A ⊕ C.<br />

2) When ϕ : A → B is a ∗-homomorphism between C ∗ -<strong>algebras</strong>, then ϕ + (a, λ) =<br />

(ϕ(a), λ) is a unital ∗-homomorphism ϕ + : A + → B + such that χB ◦ ϕ + =<br />

χA.<br />

Proof. The C ∗ -norm on A + can be obtained as follows. Define a ∗-homomorphism<br />

ψ : A + → M(A ⊕ C) by<br />

ψ(a, λ)(b, µ) = (ab + λb, λµ).<br />

It is straightforward to check that ψ is an injective ∗-homomorphism, so we obtain<br />

a norm on A + by setting x = ψ(x), x ∈ A + . We leave it to the reader to check<br />

the completeness of this norm on A + . The rest of the statements in the lemma are<br />

straightforward to verify. <br />

We are now ready for the definition of the genuine K0-group.<br />

Definition 3.18. Let A be a C ∗ -algebra. Set<br />

K0(A) = ker (χA∗ : K00(A + ) → K00(C)).<br />

For every projection e ∈ Mn(A) ⊆ Mn(A + ) we have that χA(e) = 0 so it follows<br />

that the inclusion ιA : A → A + defines us a homomorphism ιA∗ : K00(A) → K0(A).<br />

In particular, we can set<br />

K0(A) + = ιA∗(K00(A) + )<br />

to make K0(A) into a pre-ordered Abelian group. Alternatively,<br />

K0(A) + = {[e] : e ∈ P(A)}.<br />

If ψ : A → B is a ∗-homomorphism we have a unital ∗-homomorphism ψ + :<br />

A + → B + such that χB ◦ψ + = χA. It follows that ψ + ∗ : K00(A + ) → K00(B + ) maps<br />

K0(A) = ker χA∗ into ker χB∗ = K0(B). We denote the restriction<br />

ψ + ∗|K0(A) : K0(A) → K0(B)<br />

by ψ∗ and hope that we will be able to avoid any confusion with the map ψ∗ :<br />

K00(A) → K00(B). It is straightforward to check that ψ∗ : K0(A) → K0(B)<br />

is a positive homomorphism, i.e. satisfies that ψ∗(K0(A) + ) ⊆ K0(B) + and that<br />

(ϕ ◦ ψ) ∗ = ϕ∗ ◦ ψ∗, when ϕ : B → C is a ∗-homomorphism. In particular it follows<br />

that two C ∗ -<strong>algebras</strong> A and B can only be isomorphic if the groups K0(A) and<br />

K0(B) are isomorphic as pre-ordered Abelian groups. However, at this point it is<br />

not clear why K0(A) is any better than K00(A) for the purpose of distinguishing<br />

C ∗ -<strong>algebras</strong>. This question becomes even more pertinent by the next results which<br />

shows that for a large class of C ∗ -<strong>algebras</strong> these two groups agree. We will return<br />

to this issue later.<br />

Lemma 3.19. Let A be a unital C ∗ -algebra. The map ιA∗ : K00(A) → K0(A) is<br />

an isomorphism of pre-ordered Abelian groups.<br />

Proof. Let α : A + → A ⊕ C be the isomorphism from Lemma 3.17 1) and let<br />

β : K00(A ⊕ C) → K00(A) ⊕ K00(C) be the group isomorphism from Lemma 3.10.<br />

Remember that β(·) = (p1∗(·), p2∗(·)) where p1 : A⊕C → A and p2 : A⊕C → C are<br />

the two projections. Then β ◦ α∗ : K00(A + ) → K00(A) ⊕ K00(C) is an isomorphism<br />

such that β ◦ α∗ ◦ ιA∗(x) = (p1∗ ◦ α∗ ◦ ιA∗(x), p2∗ ◦ α∗ ◦ ιA∗(x)) = (idA∗(x), 0) =


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 71<br />

(x, 0), x ∈ K00(A). This relation shows that ιA∗ is injective. Let x ∈ K0(A). Since<br />

p2 ◦ α = χA and χA∗(x) = 0 we find that β ◦ α∗(x) = (y, 0) for some y ∈ K00(A).<br />

Then<br />

β ◦ α∗ ◦ ιA∗(y) = (idA∗(y), 0) = (y, 0) = β ◦ α∗(x),<br />

proving that x = ιA∗(y) because β ◦ α∗ is injective. Thus ιA∗ maps onto K0(A)<br />

and since K0(A) + = ιA∗(K00(A) + ) by definition, ιA∗ must be an isomorphism of<br />

pre-ordered Abelian groups. <br />

Lemma 3.20. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of C ∗ -<strong>algebras</strong> and A = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra. Consider the unital sequence<br />

A + 1<br />

and let ψ∞,n : A + n → lim<br />

then a ∗-isomorphism η : lim<br />

ϕ +<br />

1 <br />

A + 2<br />

ϕ +<br />

2 <br />

A + 3<br />

ϕ +<br />

3 <br />

. . .<br />

) be the corresponding canonical maps. There is<br />

for all n.<br />

−→ (A+ n , ϕ+ n<br />

(A<br />

−→ + n , ϕ+ n ) → A+ such that η ◦ ψ∞,n = ϕ + ∞,n<br />

Proof. The existence of the ∗-homomorphism η follows from the universal property<br />

of the inductive limit construction. To show that η is injective, let a ∈<br />

ker η, and let ǫ > 0. There is an n ∈ N and an element b ∈ A + n such that<br />

ψ∞,n(b) − a ≤ ǫ. Then ϕ + ∞,n (b) = η(ψ∞,n(b) − a) ≤ ǫ. Write b = (c, λ),<br />

where c ∈ An, λ ∈ C. Then ϕ + ∞,n (b) = (ϕ∞,n(c), λ) and |λ| ≤ ϕ + ∞,n (b) ≤ ǫ.<br />

Hence ϕ∞,n(c) = ϕ + ∞,n (c, 0) ≤ ϕ+ ∞,n (b) + |λ| ≤ 2ǫ. It follows ϕm,n(c) < 3ǫ<br />

for some m ≥ n, and hence ϕ + m,n (b) = (ϕm,n(c), λ) ≤ ϕm,n(c) + |λ| ≤ 4ǫ. It<br />

follows that ψ∞,n(b) ≤ 4ǫ and hence that a ≤ 5ǫ. Since ǫ > 0 was arbitrary we<br />

conclude that a = 0, and that η is injective. Let now a ∈ A + . There is then an n ∈ N,<br />

an element b ∈ An and a complex number λ such that (ϕ∞,n(b), λ) − a ≤ ǫ. Since<br />

η(ψ∞,n(b, λ)) = (ϕ∞,n(b), λ), we see that η has dense range in A + and therefore<br />

must be surjective. <br />

Theorem 3.21. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of C ∗ -<strong>algebras</strong> and A = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra. Let<br />

K0(A1) ϕ1∗ <br />

K0(A2) ϕ2∗ <br />

K0(A3) ϕ3∗ <br />

. . .<br />

be the sequence of pre-ordered Abelian groups obtained by applying the K0-functor<br />

and let G = lim<br />

−→ (K0(An), ϕn∗) be the inductive limit in the category of pre-ordered<br />

Abelian groups and q ′ n : K0(An) → G the canonical maps. There is then a unique<br />

group homomorphism q ′ : G → K0(A) such that q ′ ◦ q ′ n = ϕ∞,n ∗ for all n and q ′ is<br />

an isomorphism of pre-ordered Abelian groups.<br />

Proof. Let η : lim<br />

−→ (A + n , ϕ+ n ) → A+ be the ∗-isomorphism from Lemma 3.20. By<br />

Theorem 3.16, there is an isomorphism<br />

q : lim<br />

−→ (K00(A + n), (ϕ + n)∗) → K00(lim<br />

−→ (A + n,ϕ + n))


72 1. FUNDAMENTALS<br />

of pre-ordered groups such that q ◦ qn = ψ∞,n ∗ . Furthermore, since we have a<br />

commuting diagram<br />

K0(A1)<br />

⏐<br />

∩<br />

K00(A + 1 )<br />

ϕ1∗<br />

−−−→ K0(A2)<br />

⏐<br />

∩<br />

(ϕ +<br />

1 )∗<br />

−−−→ K00(A + 2 )<br />

ϕ2∗<br />

−−−→ K0(A3)<br />

⏐<br />

∩<br />

(ϕ +<br />

2 )∗<br />

−−−→ K00(A + 3 )<br />

ϕ3∗<br />

−−−→ . . .<br />

(ϕ +<br />

3 )∗<br />

−−−→ . . .,<br />

there is an injective map J : lim(K0(An),<br />

ϕn∗) → lim(K00(A<br />

−→ −→ + n ), (ϕ+ n )∗) such that<br />

J ◦q ′ n = qn|K0(An). Thus η∗ ◦q ◦J : lim(K0(An),<br />

ϕn∗) → K00(A<br />

−→ + ) is injective. This is<br />

in fact the map q ′ of the theorem, but to prove this, and the other assertions about<br />

it, we must first introduce more maps.<br />

Note that, because χAn+1 ◦ ϕ + n = χAn, we have a unique ∗-homomorphism χ :<br />

lim<br />

−→ (A+ n , ϕ+ n ) → C such that χ ◦ψ∞,n = χAn. Since χA ◦η ◦ψ∞,n = χA ◦ϕ∞,n + = χAn<br />

for all n, we see that χA ◦ η = χ. Similarly, since χAn+1∗ ◦ (ϕ+ n)∗ = χAn ∗ for all<br />

n, there is a unique group homomorphism χ ′ : lim(K00(A<br />

−→ + n), (ϕ + n)∗) → K00(C) such<br />

, we see that<br />

that χ ′ ◦ qn = χAn ∗ for all n. Since χ∗ ◦ q ◦ qn = χ∗ ◦ ψ∞,n∗ = χAn ∗<br />

χ∗ ◦ q = χ ′ . All in all we have that<br />

χA∗ ◦ η∗ ◦ q = χ ′ .<br />

Now, it is clear that<br />

J(lim(K0(An),<br />

ϕn∗)) = ker χ<br />

−→ ′ .<br />

(If this is not clear to the reader, he or she should provide the missing details.) So<br />

we conclude that<br />

η∗ ◦ q ◦ J(lim<br />

−→ (K0(An), ϕn∗)) ⊆ ker χA∗ = K0(A).<br />

On the other hand, if x ∈ K0(A) =<br />

ker χA∗, then x = η∗ ◦ q(y) for some y ∈ lim<br />

−→ (K00(A + n), ϕ + n ∗ ) since η∗ ◦ q is an isomorphism,<br />

and because χ ′ (y) = χA∗ ◦ η∗ ◦ q(y) = 0 we see that y ∈<br />

ker χ ′ = J(lim(K0(An),<br />

ϕn∗)). Thus η∗◦q◦J is a group isomorphism, lim(K0(An),<br />

ϕn∗) ≃<br />

−→ −→<br />

K0(A). Note that η∗◦q◦J◦q ′ n = η∗◦q◦qn|K0(An) = η∗◦ψ∞,n ∗ |K0(An) = (ϕ + ∞,n)∗|K0(An) =<br />

ϕ∞,n ∗ , so that η∗ ◦ q ◦ J = q ′ . In particular,<br />

q ′<br />

q ′ n (K0(An) + ) ⊆ K0(A) + .<br />

n<br />

If z ∈ K0(A) + , then z = η∗ ◦ q(y) for some y ∈ qn(K00(A + n )+ ) since η∗ ◦ q is an<br />

isomorphism of pre-ordered Abelian groups. Then y = qn([f]) for some f ∈ P(A + n )<br />

and<br />

[χAn(f)] = χAn∗ ([f]) = χ′ ◦ qn([f]) = χ ′ (y) = χA∗ ◦ η∗ ◦ q(y) = χA∗(z) = 0<br />

in K00(C). By definition of K00 this means that there is a projection r ∈ Mm(C)<br />

for some m such that χAn(f) ⊕ r ≈ r in Mk(C) for some k ≥ m. It follows that the<br />

trace of χAn(f) is zero and hence χAn(f) must be the zero projection. This means<br />

that f ∈ P(An) and [f] ∈ K0(An) + . Thus<br />

K0(A) + = q ′<br />

q ′ n(K0(An) + ) <br />

and q ′ is an isomorphism of pre-ordered Abelian groups.<br />

n


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 73<br />

It follows from Exercise 3.49 that K00(A) = K0(A) for a large class of C ∗ -<br />

<strong>algebras</strong>, including for example the AF-<strong>algebras</strong>. However,as will be shown in the<br />

example below,<br />

K0(C0(R 2 )) ≃ Z<br />

while<br />

K00(C0(R 2 )) = 0.<br />

So in general the two functors K00 and K0 are different and it is not generally the<br />

case that K0(A) + − K0(A) + = K0(A), although this holds for all inductive limits of<br />

unital C ∗ -<strong>algebras</strong> by Exercise 3.49 and Lemma 3.8 1). The main reason for working<br />

with K0 instead of K00 is that the former has better functorial properties.<br />

Example 3.22. This example introduces an important projection, called the<br />

Bott projection. The point is to show that projections in Mn(A) in general can not<br />

be ’diagonalized’. More precisely, we give an example of a unital C ∗ -algebra A which<br />

only contains the projections 0 and 1, while M2(A) contains a projection which is<br />

not Murray-von Neumann equivalent to any projection of the form ( a b ) for some<br />

a, b ∈ A.<br />

Let D = {z ∈ C : |z| ≤ 1} which is a compact Hausdorff space in the topology<br />

inherited from C. Thus B = C(D) is an abelian C ∗ -algebra, and since D is connected<br />

there are only the trivial projections 0 and 1 in C(D). Let T be the unit circle in C<br />

which is the topological boundary of D in C. Set<br />

A = {f ∈ B : f is constant on T}.<br />

Then A is a unital <strong>C∗</strong>-subalgebra of B which just like B only contains the projections<br />

0 and 1. Set<br />

<br />

2 1 − |z|<br />

P =<br />

z 1 − |z| 2<br />

z 1 − |z| 2 |z| 2<br />

<br />

.<br />

Since the functions f11(z) = 1 − |z| 2 , f12(z) = z 1 − |z| 2 , f21(z) = z 1 − |z| 2<br />

and f22(z) = |z| 2 all are constant on T we see that P ∈ M2(A). A straightforward<br />

check shows that P is a projection. We claim that there are no partial isometry<br />

V ∈ M2(A) such that V ∗ V = P and<br />

V V ∗ ∈ {( a b ) : a, b ∈ A} .<br />

To see this, assume that such a V exists. Since V V ∗ is a projection and A only<br />

contains the trivial projections we see that<br />

Since that<br />

V V ∗ ∈ {( 1 0 ) , ( 0 1 ) , ( 1 1 ), ( 0 0 )} .<br />

<br />

2 1 − |z| z<br />

Tr<br />

1 − |z| 2<br />

z 1 − |z| 2 |z| 2<br />

<br />

= 1,<br />

we conclude that the only possibilities are V V ∗ = ( 1 0 ) or V V ∗ = ( 0 1 ). In the last<br />

case we see that<br />

W = ( 0 1<br />

1 0 ) V<br />

will then be an element of M2(A) such that WW ∗ = ( 1 0 ) and W ∗W = P. We may<br />

therefore assume, without loss of generality, that<br />

V V ∗ = ( 1 0 ).


74 1. FUNDAMENTALS<br />

Now set<br />

S =<br />

<br />

1 − |z| 2 z<br />

,<br />

0 0<br />

which is an element of M2(B), but not of M2(A). A straigtforward check shows that<br />

It follows therefore that<br />

and<br />

Consequently<br />

S ∗ S = P, SS ∗ = ( 1 0 ).<br />

V S ∗ ( 1 0 ) = V S∗ SS ∗ = V S ∗<br />

( 1 0 ) V S∗ = V V ∗ V S ∗ = V S ∗ .<br />

V S ∗ = ( g<br />

0 )<br />

for some function g : D → C. Since<br />

<br />

|g| 2 <br />

∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗<br />

= V S (V S ) = V S SV = V PV = V V V V = V V = ( 1<br />

0 ) ,<br />

0<br />

we conclude that g takes values in T. Note that<br />

It follows that<br />

V = V V ∗ V = V S ∗ S = ( g<br />

0)S =<br />

g(z) 1 − |z| 2 g(z)z<br />

0 0<br />

<br />

.<br />

g(z)z = v12(z) (3.2)<br />

for all z ∈ D, where v12 ∈ A is the 1, 2-entry in V . Since v12 ∈ A there is complex<br />

number λ ∈ C such that w12(z) = λ for all z ∈ T. It follows from (3.2) that λ ∈ T,<br />

and then that<br />

z = λg(z) (3.3)<br />

for all z ∈ T. Define h : D → T such that h(z) = λg(z). Then h is clearly a<br />

continuous function h : D → T and (3.3) tells us that h is the identity map when<br />

restricted to the circle T. This is impossible; there are no continuous function D → T<br />

extending the identity on T. If this well-known topological fact is not familiar to<br />

the reader we offer in the following a C ∗ -algebraic proof; the involved lemma will be<br />

useful when we start to study the unitary groups of C ∗ -<strong>algebras</strong> anyway.<br />

Lemma 3.23. Let D be a unital C ∗ -algebra and u ∈ D a unitary such that<br />

1 − u < 2. It follows that there is self-adjoint element a ∈ D such that a < π<br />

and u = e ia .<br />

Proof. Note that σ(u) can not contain −1 because this would imply that the<br />

spectral radius of 1 − u is at least 2. Since the spectral radius of 1 − u is 1 − u,<br />

this is impossible by assumption. On the other hand, the spectrum of u is contained<br />

in the circle T since u is unitary so we see that<br />

σ(u) ⊆ T\{−1}.<br />

Let ϕ : T\{−1} →] − π, π[ be the inverse of ] − π, π[∋ x ↦→ e ix , and set a = ϕ(u).<br />

Then σ(a) is a compact subset of ] − π, π[ so a is self-adjoint and a < π. Since<br />

e ia = u, the proof is complete.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 75<br />

Let us use this lemma to show that there is no continuous function h : D → T<br />

such that h(z) = z for all z ∈ T. Assume that h is such a function. We can then<br />

define, for each s ∈ [0, 1], a unitary hs ∈ C(T) such that<br />

hs(z) = h(sz).<br />

Observe that the map [0, 1] ∋ s → hs ∈ C(T) is continuous; this follows from the<br />

uniform continuity of h. There is therefore a partition 0 = x0 < x1 < x2 < · · · <<br />

xN = 1 such that<br />

<br />

hxi − hxi−1<br />

≤ 1<br />

for all i = 1, 2, . . ., N. It follows then from Lemma 3.23 that there are continuous<br />

functions gj : T → [−π, π] such that<br />

hxj−1 hxj = eigj (3.4)<br />

for all j = 1, 2, . . ., N. Note that h0 is a constant; say h0 = e ir for some r ∈ R. It<br />

follows then from (3.4) that<br />

h = hxN = eiH0 ,<br />

where H0(z) = r + N<br />

j=1 gj(z). Since h(z) = z for all z ∈ T, we conclude that<br />

z = e iH0(z) for all z ∈ T or, alternatiely, that<br />

e 2πit = e 2πiH(t)<br />

for all t ∈ [0, 1], where H : [0, 1] → R is the continuous function<br />

H(t) = 1<br />

2π H0<br />

<br />

2πit<br />

e .<br />

It follows that e 2πi(t−H(t)) = 1 for all t ∈ [0, 1], and hence that t − H(t) ∈ Z for all<br />

t ∈ [0, 1]. Thus t−H(t) must be constant, which is clearly absurd since H(0) = H(1).<br />

This contradiction proves the claim.<br />

Note that A ≃ E + , where E = {f ∈ C(D) : f|T = 0}, and that<br />

[P] − [( 0 1 )] ∈ K0 (E) .<br />

To show that [P] − [( 1 0 )] is not zero in K0(E) we use the following lemma which is<br />

often usefull.<br />

Lemma 3.24. Let B be a unital C ∗ -algebra and p, q ∈ Mn(B) projections such<br />

that [p] −[q] = 0 in K0(B). There is then an m ∈ N such that p ⊕1m is Murray-von<br />

Neumann equivalent to q ⊕ 1m, where 1m = diag(1, 1, . . ., 1) ∈ Mm(B).<br />

Proof. Since [p] − [q] = 0 in K00(B) there is a projection r ∈ Mm(B), for some<br />

m, such that p ⊕ r is Murray-von Neumann equivalent to q ⊕ r. The lemma follows<br />

because r ⊕ (1m − p) is Murray-von Neumann equivalent to 1m. <br />

Assume to get a contraduction that [P] − [( 1 0 )] = 0 in K0(E). It follows from<br />

Lemma 3.24 that P ⊕1m is Murray-von Neumann equivalent to diag(1, 0, 1, 1, . . ., 1)<br />

in Mm+2(A) for some m. Thus there is a partial isometry W in Mm+2(A) such<br />

that WW ∗ = diag(1, 0, 1, 1, . . ., 1) and W ∗ W = diag(P, 1m). We compare now<br />

W with T = diag(S, 1m) which also satisfies that TT ∗ = diag(1, 0, 1, 1, . . ., 1) and<br />

T ∗ T = diag(P, 1m). Since<br />

diag(1, 0, 1, 1, . . ., 1)WT ∗ = WT ∗ diag(1, 0, 1, 1, . . ., 1) = WT ∗ ,


76 1. FUNDAMENTALS<br />

we see that the second row and second column in WT ∗ consists entirely of zeroes.<br />

Let<br />

⎛<br />

⎜<br />

X = ⎜<br />

⎝<br />

x11<br />

x21<br />

.<br />

x12<br />

x22<br />

.<br />

. . .<br />

. . .<br />

.. .<br />

⎞<br />

x1m+1<br />

x2,m+1 ⎟<br />

.<br />

⎠<br />

xm+1,1 xm+1,2 . . . xm+1,m+1<br />

be the element of Mm+1(B) obtained by removing the second row and second column<br />

in WT ∗ . Then X is a unitary because WT ∗ TW ∗ = diag(1, 0, 1, 1, . . ., 1). Since<br />

W = WT ∗ T we find that the elememt of Mm+1(A) obtained by removing the first<br />

column and the second row of W is<br />

⎛<br />

⎞<br />

zx11 x12 . . . x1m+1<br />

⎜ zx21 x22 . . . x2,m+1 ⎟<br />

⎜<br />

⎝<br />

.<br />

⎟<br />

. . .. .<br />

⎠ .<br />

zxm+1,1 xm+1,2 . . . xm+1,m+1<br />

Thus by taking determinants we obtain an element v ∈ A such that<br />

v(z) = zg(z)<br />

for all z ∈ D, where g(z) is the determinant of X(z). Since X is unitary, |g(z)| = 1,<br />

and we can obtain a contradiction in the same way as before.<br />

Thus we see that K0 (E) = 0. It is easy to see that Mn(E) contains no non-zero<br />

projections for any any n, which implies that K00 (E) = 0. Hence E is a C ∗ -algebra<br />

for which K00 (E) = K0 (E). It is easy to see that E ≃ C0 (R 2 ).<br />

3.2. The classification of AF-<strong>algebras</strong>. In this section we use the K0-group<br />

to classify AF-<strong>algebras</strong> up to isomorphism.<br />

In the following A is a C ∗ -algebra.<br />

Definition 3.25. A has cancellation of projections when the following holds:<br />

When n ∈ N and e, f ∈ Mn(A) are projections such that [e] = [f] in K0(A), it<br />

follows that e ≈ f, i.e. that e and f are Murray-von Neumann equivalent in Mn(A).<br />

Lemma 3.26. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

be a sequence of C ∗ -<strong>algebras</strong> and A = lim<br />

−→ (An, ϕn) the corresponding inductive limit<br />

C ∗ -algebra. Assume that each Ai has cancellation of projections. Then A has cancellation<br />

of projections.<br />

Proof. Let e, f ∈ Mn(A) be projections such that [e] = [f] in K0(A). By<br />

Lemma 3.15 there is then an m ∈ N and projections e ′ , f ′ ∈ Mn(Am) such that<br />

ϕ∞,m(e ′ ) − e ≤ 1<br />

4 and ϕ∞,m(f ′ ) − f ≤ 1<br />

. It follows from Lemma 3.14 that<br />

4<br />

e ≈ ϕ∞,m(e ′ ) and f ≈ ϕ∞,m(f ′ ). In particular, ϕ∞,m∗ [e ′ ] = [e] = [f] = ϕ∞,m∗ [f ′ ].<br />

And then Theorem 3.21 shows that ϕk,m∗ [e ′ ] = ϕ∞,m∗ [f ′ ] in K0 (Ak) for some k ><br />

m. Since Ak has cancellation of projections it follows that ϕk,m(e ′ ) ≈ ϕk,m(f ′ ) in<br />

Mn (Ak). Consequently, ϕ∞,m(e ′ ) ≈ ϕ∞,m(f ′ ) in Mn (A). <br />

Lemma 3.27. Let A be a <strong>C∗</strong>-algebra such that  has cancellation of projections.<br />

1) When e, f ∈ P(A) are two projections, we have that [e] = [f] in K0(A) if<br />

and only if e ≈ f.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 77<br />

2) When e and f are projections in A such that [e] = [f] in K0(A), there is a<br />

unitary u ∈ Â such that ueu∗ = f.<br />

3) K0(A) is a partially ordered group, i.e. K0(A) + ∩ (−K0(A) + ) = {0}.<br />

Proof. 1) Assume that [e] = [f] in K0(A). Then [e] = [f] in K0( Â) and hence<br />

e ≈ f since  has cancellation of projections. The converse is trivial.<br />

2) : From 1) we get a partial isometry v ∈ A such that vv ∗ = f, v ∗ v = e. Since<br />

[1 −e]+[e] = [1] = [1 −f]+[f] in K0( Â) by Lemma 3.2 we see that [1 −e] = [1 −f]<br />

in K0( Â). Since  has cancellation of projections we conclude that 1 − f ≈ 1 − e<br />

in Â, i.e. there is a partial isometry w ∈ Â such that ww∗ = 1 − f, w∗w = 1 − e.<br />

Then u = v + w is a unitary in  such that ueu∗ = f.<br />

3) : Let x ∈ K0(A) + ∩ (−K0(A) + ). There are projections e, f ∈ P(A) such that<br />

x = [e] = −[f] in K0(A). But this means that [e ⊕ f] = 0 in K0(A). By 1) this<br />

implies that e ⊕ f ≈ 0, and hence that e ⊕ f = 0. Thus e = f = 0 and x = 0 in<br />

K0(A).<br />

<br />

Let A be a C ∗ -algebra. A continuous linear functional ω ∈ A ∗ is a trace when<br />

ω(xy) = ω(yx), x, y ∈ A.<br />

Lemma 3.28. Let ω ∈ A∗ be a trace. There is then a trace ωn ∈ Mn(A) ∗ such<br />

that<br />

<br />

ωn (aij) n<br />

n<br />

i,j=1 = ω (aii).<br />

Moreover, every trace of Mn(A) is of this form. Specifically if µ ∈ Mn(A) ∗ is a trace,<br />

⎛ ⎞<br />

a 0 0 . . . 0<br />

⎜0<br />

0 0 . . . 0⎟<br />

⎜ ⎟<br />

ω(a) = µ ⎜0<br />

0 0 . . . 0⎟<br />

⎜<br />

⎝ .<br />

⎟<br />

. . . .. . ⎠<br />

0 0 0 . . . 0<br />

defines a trace ω ∈ A ∗ such that µ = ωn.<br />

Proof. Note that<br />

<br />

<br />

ωn<br />

≤<br />

<br />

(aij) n<br />

i,j=1<br />

i=1<br />

n<br />

|ω (aii)| ≤ nω max<br />

i aii<br />

<br />

<br />

≤ nω<br />

i=1<br />

This shows that ωn ∈ Mn(A) ∗ . When A = (aij) n<br />

i,j=1<br />

ωn (AB) =<br />

=<br />

n<br />

<br />

n<br />

ω<br />

i=1<br />

k=1<br />

aikbki<br />

n<br />

ω (bikaki) = ωn(BA).<br />

i,k=1<br />

<br />

=<br />

n<br />

ω (aikbki) =<br />

i,k=1<br />

(aij) n<br />

i,j=1<br />

<br />

<br />

.<br />

n<br />

and B = (bij)<br />

i,j=1 , we find that<br />

n<br />

ω (bkiaik)<br />

We leave the proof of the remaining part of the statement to the reader. <br />

Definition 3.29. A C ∗ -algebra A is rich on traces when the following holds:<br />

When n ∈ N and e, f ∈ Mn(A) are projections such that ω(e) = ω(f) for all traces<br />

ω ∈ Mn(A) ∗ , it follows that e ≈ f.<br />

i,k=1


78 1. FUNDAMENTALS<br />

Lemma 3.30. A C ∗ -algebra A which is rich on traces has cancellation of projections.<br />

Proof. Let n ∈ N and assume that e, f ∈ Mn(A) are projections such that<br />

[e] = [f] in K0(A). There is then a projection p ∈ Mm(A + ) such that e ⊕ p ≈ f ⊕ p<br />

in Mn+m(A + ). Let µ ∈ Mn(A) ∗ be a trace. By Lemma 3.28 there is a trace ω ∈ A ∗<br />

such that µ = ωn. Define ˜ω : A + → C such that ˜ω(a, λ) = ω(a) + λ and note that<br />

˜ω ∈ A ∗ is a trace. Set ˜µ = ˜ωn+m ∈ Mn+m(A + ) ∗ which is a trace on Mn+m(A + ).<br />

Since e ⊕ p ≈ f ⊕ p we find that<br />

˜µ(e ⊕ p) = ˜µ(f ⊕ p).<br />

Note that ˜µ(e ⊕ p) = ˜ωn(e) + ˜ωm(p) = ωn(e) + ˜ωm(p) = µ(e) + ˜ωm(p) and similarly<br />

˜µ(f ⊕ p) = µ(f) + ˜ωm(p). Hence µ(e) = µ(f). Since µ was an arbitrary trace in<br />

Mn(A) ∗ it follows that e ≈ f since A is rich on traces. <br />

Lemma 3.31. A finite dimensional C ∗ -algebra is rich on traces.<br />

Proof. Let F be a finite-dimensinal <strong>C∗</strong>-algebra and e, f ∈ F projections such<br />

that ω(e) = ω(f) for every trace ω ∈ A∗ . To show that e ≈ f we may assume<br />

that F = Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN , cf. Theorem 1.60. Then we can write e =<br />

(e1, e2, . . .,eN), f = (f1, f2, . . .,fN) where ei, fi ∈ Mni . For each i ∈ {1, 2, . . ., N}<br />

we can define a trace Tri : F → C by<br />

Tri(x1, x2, . . .,xN) = Tr(xi),<br />

(x1, . . .,xN) ∈ F. Since Tr(ei) = Tri(e) = Tri(f) = Tr(fi) by assumption, we know<br />

from linear algebra that there is a unitary ui ∈ Mni such that uieiui ∗ = fi, i =<br />

1, 2, . . ., N. Set u = (u1, u2, . . ., uN) and note that ueu ∗ = f.<br />

Since Mn(F) is a finite dimensional C ∗ -algebra for all n, this shows that F is<br />

rich on traces. <br />

Lemma 3.32. Let A be an AF-algebra.<br />

1) When e, f ∈ P(A) are two projections, we have that [e] = [f] in K0(A) if<br />

and only if e ≈ f.<br />

2) When e and f are projections in A such that [e] = [f] in K0(A), there is a<br />

unitary u ∈ Â such that ueu∗ = f.<br />

3) K0(A) is a partially ordered group, i.e. K0(A) + ∩ (−K0(A) + ) = {0}.<br />

Proof. It follows from Lemma 3.20 that  is the inductive limit of a sequence<br />

of finite-dimensional <strong>C∗</strong>-<strong>algebras</strong> (with unital connecting maps). Therefore  has<br />

cancellation of projections by Lemma 3.31. Hence Lemma 3.27 applies. <br />

Definition 3.33. For any C ∗ -algebra A we set<br />

Σ(A) = {[e] ∈ K0(A) : e ∈ P1(A)}.<br />

Thus Σ(A), called the scale of A, consists of the elements in K0(A) which are<br />

represented by projections in A. Note that Σ(A) ⊆ K0(A) + .<br />

Lemma 3.34. Let A be a finite dimensional C ∗ -algebra and B = lim<br />

−→ (Bn, ψn) the<br />

inductive limit of a sequence<br />

B1<br />

ψ1 <br />

B2<br />

of finite direct sums of matrix <strong>algebras</strong>.<br />

ψ2 <br />

B3<br />

ψ3 <br />

. . .


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 79<br />

1) If r : K0(A) → K0(B) is a positive homomorphism such that r(Σ(A)) ⊆<br />

Σ(B), then there is a ∗-homomorphism ϕ : A → Bk for some k ∈ N such<br />

that ψ∞,k ∗ ◦ ϕ∗ = r.<br />

2) If µ, λ : A → B are two ∗-homomorphisms such that µ∗ = λ∗ on K0(A),<br />

then there is a unitary u ∈ ˆ B such that<br />

Ad u ◦ λ = µ.<br />

Proof. 1) : Before reading the proof the reader is advised to do Exercise 3.50<br />

} be a complete set of matrix units in A. Since r(Σ(A)) ⊆ Σ(B) there<br />

first. Let {ed i,j<br />

are projections qd i ∈ B such that<br />

for all d, i. Furthermore, since <br />

d,i<br />

there is a projection p ∈ B such that<br />

r([e d ii]) = [q d i ]<br />

[q d i ] = r([1]) ∈ Σ(B)<br />

[diag q 1 1 , . . .,q1 n1 , q2 1 , . . .,q2 n2 , q3 1 , . . ....,qN nN<br />

] = <br />

d,i<br />

[q d i<br />

] = [p]<br />

in K0(B). By Theorem 3.21 or Theorem 3.16 there is a k0 ∈ N and projections<br />

hd i ∈ Bk0 and p ′ ∈ Bk0 such that qd <br />

d<br />

i = ψ∞,k0∗ hi and [p] = ψ∞,k0∗ [p′ ] in K0(B)<br />

and<br />

<br />

] =<br />

[diag h 1 1 , . . .,h1 n1 , h2 1 , . . .,h2 n2 , h3 1 , . . ....,hN nN<br />

d,i<br />

[h d i ] = [p′ ]<br />

in K0(Bk0). It follows that there is a partial isometry V ∈ MPN d=1<br />

nd (Bk) such that<br />

⎛<br />

⎞<br />

h 1 1 0 0 . . . 0<br />

V ∗ ⎜ 0<br />

⎜<br />

V = ⎜<br />

⎝<br />

h1 0<br />

2 0<br />

h<br />

. . . 0<br />

1 . .<br />

3<br />

.<br />

. . .<br />

. ..<br />

0<br />

.<br />

0 0 0 . . . hN ⎟<br />

nN<br />

and<br />

⎠ = diag h 1 1 , . . .,h1 n1 , h2 1 , . . ., h2 n2 , h3 1 , . . ....,hN nN<br />

V V ∗ ⎛<br />

p<br />

⎜<br />

= ⎜<br />

⎝<br />

′ ⎞<br />

0 0 . . . 0<br />

0 0 0 . . . 0 ⎟<br />

0 0 . . . 0⎟<br />

.<br />

⎟<br />

. . . .. . ⎠<br />

0 0 0 . . . 0<br />

.<br />

There are then projections s d i ∈ Bk such that<br />

⎛<br />

s d i 0 0 . . . 0<br />

⎞<br />

⎜ 0 0 0 . . . 0⎟<br />

⎜<br />

⎟<br />

⎜ 0 0 . . . 0⎟<br />

⎜<br />

⎝ ..<br />

⎟<br />

. . . . . ⎠<br />

0 0 0 . . . 0<br />

= V diag 0, 0, . . ., 0, h d i , 0, . . .,0)V ∗ .<br />

Note that the sd i ’s are mutually orthogonal: sdi sd′<br />

i ′ = 0 unless d = d′ and i = i ′ . Note<br />

d d d d d d that ψ∞,k0∗ si = ψ∞,k0 ∗ hi = qi = r eii = r e11 = ψ∞,k0∗ s1 for all d, i.<br />

By Theorem 3.21 (or Theorem 3.16) this implies that there is a k > k0 such that


80 1. FUNDAMENTALS<br />

<br />

d d si = ψk,k0 s1 in K0 <br />

(Bk)<br />

<br />

for all d, i. There are therfore partial isometries<br />

d i s1 and vdvi ∗ <br />

d<br />

d = ψk,k0 si for all i, d. By using that<br />

’s are mutually orthogonal it is the easy to check that<br />

<br />

ψk,k0<br />

vi d ∈ Bk such that vi ∗ i<br />

d vd = ψk,k0<br />

the sd i<br />

f d ij = v i dv j ∗<br />

d<br />

d = 1, 2, . . ., N, i, j = 1, 2, . . ., nd, is a set of matrix units in Bk. There is therefore<br />

d = fij for all i, j, d. Since<br />

a ∗-homomorphism ϕ : A → Bk such that ϕ ed ij<br />

<br />

d d d d d<br />

ψ∞,k∗ ◦ ϕ∗ e11 = ψ∞,k∗ f11 = ψ∞,k ψk,k0 ∗ si = ψ∞,k0∗ si = r e11 ,<br />

for all d, it follows that ψ∞,k∗ ◦ ϕ∗ = r as desired.<br />

2) : Since [µ(ed 11 )] = µ∗([ed 11 ]) = λ∗([ed 11 ]) = [λ(ed11 )] in K0(B) we conclude from<br />

Lemma 3.27 1) that there is a partial isometry vd ∈ B such that vdvd ∗ = µ(ed 11 ) and<br />

vd ∗ vd = λ(e d 11<br />

). Set<br />

v = <br />

µ(e d i1)vdλ(e d 1i).<br />

d,i<br />

Then vv ∗ = µ(1), v ∗ v = λ(1) and vλ(e d ij)v ∗ = µ(e d ij) for all d, i, j. Since [λ(1)] =<br />

[µ(1)] in K0(B) there is, by Lemma 3.27 2) a unitary s ∈ ˆ B such that s(1−λ(1))s ∗ =<br />

1 − µ(1). Then w = s(1 − λ(1)) ∈ ˆ B is a partial isometry such that ww ∗ =<br />

1 −µ(1), w ∗ w = 1 −λ(1) and we set u = v+w. Then u is a unitary in B + such that<br />

Adu ◦ λ = µ. (The calculations are essentially the same as in the proof of Lemma<br />

2.6.) <br />

and<br />

We can now prove the classification theorem for AF-<strong>algebras</strong>.<br />

Theorem 3.35. Let A and B be AF-<strong>algebras</strong>.<br />

1) If f : K0(A) → K0(B) is a positive group homomorphism such that f(Σ(A)) ⊆<br />

Σ(B), then there is a ∗- homomorphism ϕ : A → B such that ϕ∗ = f.<br />

2) If ϕ, ψ : A → B are two ∗-homomorphisms such that ψ∗ = ϕ∗ on K0(A),<br />

then there is a sequence un of unitaries in B + such that<br />

ϕ(a) = lim<br />

n→∞ unψ(a)u ∗ n, a ∈ A.<br />

3) When f : K0(A) → K0(B) is an isomorphism of partially ordered groups<br />

such that f(Σ(A)) = Σ(B), then there is an isomorphism ϕ : A → B such<br />

that ϕ∗ = f.<br />

Proof. 1) : Let A and B be the inductive limits of<br />

A1<br />

B1<br />

ϕ1 <br />

A2<br />

ψ1 <br />

B2<br />

ϕ2 <br />

A3<br />

ψ2 <br />

B3<br />

ϕ3 <br />

. . .<br />

ψ3 <br />

. . .<br />

respectively, where each An and Bn is a finite direct sum of matrix <strong>algebras</strong>. We will<br />

construct sequences l1 < l2 < . . . and k1 < k2 < . . . in N and ∗-homomorphisms<br />

λi : Ali → Bki such that ψ∞,ki∗ ◦ λi∗ = f ◦ ϕ∞,li∗ and the diagram<br />

λ1<br />

Al1<br />

⏐<br />

<br />

Bk1<br />

ϕl2 ,l1 −−−→ Al2<br />

λ2<br />

⏐<br />

<br />

ψk 2 ,k 1<br />

−−−→ Bk2<br />

ϕl3 ,l2 −−−→ Al3<br />

λ3<br />

⏐<br />

<br />

ψk 3 ,k 2<br />

−−−→ Bk3<br />

ϕl 4 ,l 3<br />

−−−→ . . .<br />

ψk 4 ,k 3<br />

−−−→ . . .


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 81<br />

commutes. Assume that we have constructed everything up to n ∈ N. We then<br />

construct ln+1, kn+1 and λn+1 such that<br />

λn<br />

Aln<br />

⏐<br />

<br />

Bkn<br />

ϕln+1 ,ln<br />

−−−−−→ Aln+1<br />

λn+1<br />

⏐<br />

<br />

ψk n+1 ,kn<br />

−−−−−→ Bkn+1<br />

commutes and ψ∞,kn+1∗ ◦ λn+1∗ = f ◦ ϕ∞,ln+1∗ . First we take an arbitrary ln+1 > ln.<br />

By Lemma 3.34 1) there is an m > kn and a ∗- homomorphism λ : Aln+1 → Bm<br />

such that ψ∞,m∗ ◦ λ∗ = f ◦ ϕ∞,ln+1∗ . In comparison we have that ψ∞,kn∗ ◦ λn∗ =<br />

f ◦ ϕ∞,ln∗ = f ◦ ϕ∞,ln+1∗ ◦ ϕln+1,ln ∗ . Hence ψ∞,m∗ ◦ λ∗ ◦ ϕln+1,ln∗ = ψ∞,kn∗ ◦ λn∗.<br />

Let {ed ij } be the standard matrix units in Aln. By Theorem 3.21 (or Theorem 3.16)<br />

there is a kn+1 > m such that<br />

for all d. But then<br />

ψkn+1,m ∗ ◦ λ∗ ◦ ϕln+1,ln ∗ ([ed 11]) = ψkn+1,kn∗ ◦ λn∗([e d 11])<br />

= ψkn+1,kn∗ ◦ λn∗<br />

ψkn+1,m ◦ λ∗ ◦ ϕln+1,ln ∗ ∗<br />

on K0(Aln) by Exercise 3.50. By Lemma 3.34 2) there is therefore a unitary u ∈ B +<br />

kn+1<br />

such that<br />

Ad u ◦ ψkn+1,m ◦ λ ◦ ϕln+1,ln = ψkn+1,kn ◦ λn.<br />

We set λn+1 = Ad u ◦ ψkn+1,m ◦ λ and observe that the desired diagram commutes<br />

with this choice for λn+1. Since Ad u induces the identity map on K0(Bkn+1) we find<br />

that<br />

ψ∞,kn+1∗ ◦ λn+1∗ = ψ∞,kn+1∗ ◦ ψkn+1,m ∗ ◦ λ∗ = ψ∞,m ∗ ◦ λ∗ = f ◦ ϕ∞,ln+1∗ .<br />

The diagram gives immediately the existence of a ∗-homomorphism ϕ : A → B<br />

such that<br />

ϕ ◦ ϕ∞,ln = ψ∞,kn ◦ λn, n ∈ N.<br />

Since ϕ∗ ◦ ϕ∞,ln∗ = ψ∞,kn∗ ◦ λn∗ = f ◦ ϕ∞,ln∗ for all n, we see that ϕ∗ = f.<br />

2) : Since<br />

(ϕ ◦ ϕ∞,n)∗ = ϕ∗ ◦ ϕ∞,n ∗ = ψ∗ ◦ ϕ∞,n ∗ = (ψ ◦ ϕ∞,n)∗<br />

on K0(An), Lemma 3.34 2) provides a unitary un ∈ B + such that Ad un ◦ϕ◦ϕ∞,n =<br />

ψ ◦ ϕ∞,n. This gives us a sequence of unitaries, un, and since <br />

i ϕ∞,i(Ai) = A, it is<br />

clear that limn→∞ unϕ(a)u∗ n = ψ(a), a ∈ A.<br />

3) : We will construct sequences l1 < l2 < l3 < · · · and k1 < k2 < · · · in N and ∗homomorphisms<br />

λn : Aln → Bkn, µn : Bkn → Aln+1 such that ψ∞,kn∗ ◦λn∗ = f ◦ϕ∞,ln∗<br />

and ϕ∞,ln+1∗ ◦ µn∗ = f −1 ◦ ψ∞,kn∗ for all n and<br />

λ1<br />

Al1<br />

<br />

ϕl2 ,l ϕl 1 3 ,l ϕl 2 4 ,l3 <br />

Al2<br />

<br />

Al3<br />

<br />

<br />

µ1 <br />

<br />

<br />

<br />

λ2<br />

<br />

<br />

µ2 <br />

<br />

<br />

<br />

λ3<br />

<br />

. <br />

. .<br />

<br />

µ3 <br />

<br />

<br />

<br />

<br />

. . .<br />

Bk1<br />

<br />

Bk2<br />

<br />

Bk3<br />

ψk2 ,k ψk 1 3 ,k ψk 2 4 ,k3 commutes. Once this has been obtained it is easy to finish the proof : There are<br />

∗-homomorphisms ϕ : A → B and ψ : B → A such that<br />

ϕ ◦ ϕ∞,ln = ψ∞,kn ◦ λn


82 1. FUNDAMENTALS<br />

and<br />

ψ ◦ ψ∞,kn = ϕ∞,ln+1 ◦ µn<br />

for all n and hence ϕ and ψ must be the inverse of each other. That ϕ∗ = f follows<br />

as in the proof of 1).<br />

The desired diagram is constructed by induction. Let us assume that everything<br />

has been constructed for n ≤ N, i.e. that we have everything up to<br />

ϕlN ,lN−1 AlN−1<br />

AlN<br />

λN−1<br />

<br />

<br />

<br />

<br />

<br />

<br />

µN−1<br />

<br />

λN<br />

BkN−1<br />

<br />

BkN<br />

ψkN ,kN−1 .<br />

To construct µN we repeat the argument used in the proof of 1). By Lemma 3.34 1)<br />

there is an m > lN and a ∗-homomorphism µ : BkN → Am such that ϕ∞,m∗ ◦ µ∗ =<br />

f −1 ◦ ψ∞,kN ∗ . Then<br />

ϕ∞,m ∗ ◦ (µ ◦ λN) ∗ = f −1 ◦ ψ∞,kN ∗ ◦ λN ∗ = f −1 ◦ f ◦ ϕ∞,lN ∗ = ϕ∞,lN ∗<br />

so we conclude that there is a lN+1 > m such that<br />

ϕlN+1,m ∗ ◦ (µ ◦ λN) ∗ = ϕlN+1,lN ∗ .<br />

By Lemma 3.34 2) there is then a unitary v ∈ A +<br />

such that<br />

lN+1<br />

Ad v ◦ ϕlN+1,m ◦ µ ◦ λN = ϕlN+1,lN .<br />

Set µN = Ad v ◦ ϕlN+1,m ◦ µ. Then ϕ∞,lN+1∗ ◦ µN ∗ = ϕ∞,m ∗ ◦ µ∗ = f −1 ◦ ψ∞,kN ∗ .<br />

The construction of kN+1 > kN and λN+1 : AlN+1 → BkN+1 is similar and should be<br />

routine now. <br />

The classification result for unital AF-<strong>algebras</strong> can be improved a little by use<br />

of the following lemma.<br />

Lemma 3.36. Let A be a unital AF-algebra. Then<br />

Σ(A) = {x ∈ K0(A) : 0 ≤ x ≤ [1]}.<br />

Proof. When p is a projection in A we have that<br />

[1] = [p] + [1 − p]<br />

in K0(A), cf. the proof of Lemma 3.9. Hence 0 ≤ [p] ≤ [1], proving that<br />

Σ(A) ⊆ {x ∈ K0(A) : 0 ≤ x ≤ [1]}.<br />

To prove the reverse inclusion, let x ∈ K0(A) such that 0 ≤ x ≤ [1]. Then x = [p]<br />

for some projection p ∈ Mn(A) and by Lemma 3.27 1) there is a projection r ∈<br />

Mm(A) such that p ⊕ r ≈ 1. Let v ∈ Mn+m(A) be a partial isometry such that<br />

vv ∗ = 1, v ∗ v = p ⊕r. Set q = vpv ∗ and note that q is a projection : q 2 = vpv ∗ vpv ∗ =<br />

vp(p ⊕ r)pv ∗ = vpv ∗ = q and that q = vpv ∗ ≤ v1v ∗ = 1. Hence q is a projection in<br />

A (strictly speaking, q is a projection in Mn+m(A) whose only non-zero entry is in<br />

the upper left-hand corner). Hence x = [p] = [q] ∈ Σ(A). <br />

Theorem 3.37. Let A and B be unital AF-<strong>algebras</strong>.<br />

1) If f : K0(A) → K0(B) is a positive group homomorphism such that f([1]) =<br />

[1], then there is a unital ∗- homomorphism ϕ : A → B such that ϕ∗ = f.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 83<br />

2) If ϕ, ψ : A → B are two unital ∗-homomorphisms such that ψ∗ = ϕ∗ on<br />

K0(A), then there is a sequence un of unitaries in B such that<br />

ϕ(a) = lim<br />

n→∞ unψ(a)u ∗ n, a ∈ A.<br />

3) When f : K0(A) → K0(B) is an isomorphism of partially ordered groups<br />

such that f([1]) = [1], then there is an isomorphism ϕ : A → B such that<br />

ϕ∗ = f.<br />

Proof. 1): By combining Lemma 3.36 with Theorem 3.35 1) we get a ∗homomorphism<br />

ψ : A → B such that ψ∗ = f. Since [ψ(1)] = f([1]) = [1] in K0(B)<br />

we conclude from Lemma 3.27 that there is a unitary u ∈ B such that uψ(1)u ∗ = 1.<br />

Set ϕ = Ad u ◦ ψ.<br />

2) and 3) follow straightforwardly from Lemma 3.36 and Theorem 3.35. <br />

3.3. The range of the invariant : Dimension groups. The classification<br />

result, Theorem 3.35, says that the isomorphism class of an AF-algebra is completely<br />

determined by the K0-group, considered as a partially ordered Abelian group, plus<br />

the position of the scale in this group or the position of [1] in the unital case, cf.<br />

Theorem 3.37. In this section we give an abstract description of the partially ordered<br />

groups which occur in this way, i.e. as the K0-group of an AF-algebra. We will then,<br />

subsequently, determine the subsets of the group which can play the role of the scale<br />

of an AF-algebra.<br />

We emphasize that the K0-group of an arbitrary C ∗ -algebra is not partially<br />

ordered. For an AF-algebra A, however, K0(A) is always partially ordered by Lemma<br />

3.27 3).<br />

Lemma 3.38. Let G be a partially ordered Abelian group. Then the following<br />

conditions are equivalent :<br />

1) G is upward directed, i.e. for all g, h ∈ G there is a k ∈ G such that g ≤ k<br />

and h ≤ k.<br />

2) G is downward directed, i.e. for all g, h ∈ G there is a k ∈ G such that<br />

k ≤ g and k ≤ h.<br />

3) G is generated by G + .<br />

4) G = G + − G + .<br />

Proof. 1) : ⇔ 2) follows by using the map g ↦→ −g.<br />

1) ⇒ 4) : Given a ∈ G, upward directedness implies that there is some x ∈ G<br />

such that a ≤ x and 0 ≤ x. Then a = x − (x − a) and x, x − a ∈ G + .<br />

4) ⇒ 1) : Given a1, a2 ∈ G, write ai = xi − yi for some xi, yi ∈ G + , i = 1, 2. Set<br />

a = x1 + x2 and note that a1, a2 ≤ a.<br />

3) ⇒ 4) : Let a ∈ G. Since a is in the group generated by G + , there are integers<br />

zi ∈ Z and elements gi ∈ G + , i = 1, 2, . . ., n, such that<br />

a = z1g1 + z2g2 + · · · + zngn.<br />

For each i, write zi = si − ti, where si, ti ∈ N. Then x = n<br />

i=1 sigi ∈ G + , y =<br />

n<br />

i=1 tigi ∈ G + and a = x − y. 4) ⇒ 3) is trivial. <br />

A partially ordered Abelian group is directed when the equivalent conditions of<br />

Lemma 3.38 are satisfied. Note that Exercise 3.49 shows that the K0-group of an<br />

AF-algebra is directed.


84 1. FUNDAMENTALS<br />

Proposition 3.39. Let G be a partially ordered Abelian group. Then the following<br />

conditions are equivalent :<br />

1) If x1, x2, y1, y2 ∈ G and xi ≤ yj for all i, j, then there exists z ∈ G such that<br />

xi ≤ z ≤ yj for all i, j.<br />

2) If x, y1, y2 ∈ G + and x ≤ y1 + y2, then x = x1 + x2 for some x1, x2 ∈ G +<br />

such that xj ≤ yj, j = 1, 2.<br />

3) If x1, x2, y1, y2 ∈ G + and x1 + x2 = y1 + y2, then there exist elements zij ∈<br />

G + , i, j = 1, 2, such that xi = zi1 +zi2, i = 1, 2 and yj = z1j +z2j, j = 1, 2.<br />

Proof. 1) ⇒ 2) : We are given 0 ≤ x and 0 ≤ y2. Since y1 ≥ 0, we also have<br />

x − y1 ≤ x and x − y1 ≤ y2. By 1), there is some x2 ∈ G such that 0 ≤ x2 ≤ x and<br />

x − y1 ≤ x2 ≤ y2. Set x1 = x − x2. Then x1, x2 ∈ G + with x = x1 + x2 and x2 ≤ y2.<br />

As x − y1 ≤ x2, we also have x1 = x − x2 ≤ y1.<br />

2) ⇒ 3) : Since x2 ≥ 0, we have x1 ≤ y1 + y2, so by 2), x1 = z11 + z12 for some<br />

z11, z12 ∈ G + such that z1j ≤ yj, j = 1, 2. Set z2j = yj − z1j, j = 1, 2, so that<br />

z2j ∈ G + and yj = z1j + z2j. Since<br />

x1 + x2 = y1 + y2 = z11 + z21 + z12 + z22<br />

we conclude that x2 = z21 + z22.<br />

3) ⇒ 1) : We have that yj − xi ∈ G + for all i, j and<br />

(y1 − x1) + (y2 − x2) = (y1 − x2) + (y2 − x1).<br />

By 3) there exist elements zij ∈ G + , i, j = 1, 2, such that yi −xi = zi1 +zi2, i = 1, 2<br />

while also y1 −x2 = z11 +z21 and y2 −x1 = z12 +z22. Set z = x1 +z12, and note that<br />

x1 ≤ z. As y2 −x1 = z12 +z22, we obtain z = y2 −z22 ≤ y2. On the other hand, from<br />

y1 − x1 = z11 + z12, we see that z = y1 − z11 ≤ y1. Finally, since y1 − x2 = z11 + z21,<br />

we have that x2 ≤ x2 + z21 = y1 − z11 = z. Thus xi ≤ z ≤ yj for all i, j. <br />

Property 1) of Proposition 3.39 is called the Riesz interpolation property, while<br />

properties 2),3) are called the Riesz decomposition property. A partially ordered<br />

Abelian group is an interpolation group when it has one and hence both of these<br />

properties.<br />

In the present context it is of crucial importance that the interpolation property<br />

is preserved under inductive limits.<br />

Proposition 3.40. Let<br />

G1<br />

µ1 <br />

G2<br />

µ2 <br />

G3<br />

µ3 <br />

. . .<br />

be a sequence of interpolation groups with positive connecting maps. Then the inductive<br />

limit group G = lim<br />

−→ (Gn, µn) is an interpolation group.<br />

Proof. Let qn : Gn → G denote the canonical homomorphisms. Let xi, yj ∈ G<br />

such that xi ≤ yj, i, j = 1, 2. There is a k ∈ N such that<br />

xi, yj ∈ qk(Gk)<br />

for all i, j. Take ai, bj ∈ Gk such that qk(ai) = xi, qk(bj) = yj for all i, j. Since<br />

yj − xi ∈ G + , there is an m > k such that<br />

µm,k(bj − ai) ∈ G + m<br />

for all i, j. Since Gm is an interpolation group, there is an element c ∈ Gm such that<br />

µm,k(ai) ≤ c ≤ µm,k(bj) for all i, j. Set z = qm(c). Then xi ≤ z ≤ yj for all i, j.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 85<br />

A partially ordered Abelian group G is lattice ordered when the following holds:<br />

For each pair a, b ∈ G there is an element, denoted by a ∨ b, such that<br />

and<br />

a, b ≤ a ∨ b<br />

c ∈ G, a, b ≤ c ⇒ a ∨ b ≤ c.<br />

In other words, the supremum of a and b exists.<br />

In is clear that a lattice ordered group is also an interpolation group. However,<br />

the property of being lattice ordered is not preserved in inductive limits. A<br />

prominent example of a lattice ordered group is Z n , ordered in the natural way by<br />

Z n+ = {(z1, z2, . . .,zn) ∈ Z n : zi ≥ 0, i = 1, 2, . . ., n}.<br />

A partially ordered group which is isomorphic to Z n , ordered in this way (for some<br />

n) is called a simplicial group. The K0-group of a finite-dimensional C ∗ -algebra is a<br />

simplicial group by Exercise 3.50, so we see from Theorem 3.21 that the K0-group of<br />

an AF-algebra is the inductive limit of a sequence of simplicial groups. In particular,<br />

the K0-group of an AF-algebra is an interpolation group.<br />

Now we only need one more order theoretic property in order to specify which<br />

partially ordered groups occur as K0 of an AF-algebra. A partially ordered Abelian<br />

group G is unperforated when the implication n ∈ N, ng ≥ 0 ⇒ g ≥ 0 holds for all<br />

g ∈ G. Note that Z n , ordered by N n , is unperforated and that the inductive limit of<br />

a sequence of unperforated partially ordered Abelian groups is again unperforated.<br />

Hence the K0-group of an AF-algebra is always unperforated.<br />

All in all we see that the K0-group of an AF-algebra enjoys the following properties<br />

1) directedness,<br />

2) the Riesz interpolation property,<br />

3) unperforatedness.<br />

A partial ordered Abelian group with these three properties will be called a<br />

dimension group. In other words, a dimension group is an interpolation group which<br />

is directed and unperforated. An obvious example of a dimension group is R, the<br />

real numbers, with its natural ordering. However, R is not countable, and can<br />

therefore not be the K0-group of an AF-algebra. The K0-group of an AF-algebra is<br />

the inductive limit of a sequence of countable groups and must therefore also itself<br />

be countable. This size restriction is the last restriction, and it is the only restriction<br />

which is not directly related to the partial ordering. But before we can present a<br />

proof of this assertion we need some preparations. The main problem is to show<br />

that a countable dimension group is the inductive limit of a sequence of simplicial<br />

groups.<br />

We first need the following strong form of the Riesz decomposition properties.<br />

Lemma 3.41. Let G be a dimension group and let x1, x2, . . .,xn ∈ G + and<br />

a1, a2, . . .,an ∈ Z such that<br />

a1x1 + a2x2 + · · · + anxn = 0.<br />

Then there exist elements y1, y2, . . ., ym ∈ G + and nonnegative integers bij, i =<br />

1, 2, . . ., n, j = 1, 2, . . ., m, such that<br />

xi = bi1y1 + bi2y2 + · · · + bimym


86 1. FUNDAMENTALS<br />

for all i = 1, 2, . . ., n, and<br />

for all j = 1, 2, . . ., m.<br />

a1b1j + a2b2j + · · · + anbnj = 0<br />

Proof. Let N×N have the lexicographic ordering, i.e. (a, b) ≤ (a1, b1) iff a < a1<br />

or a = a1 and b ≤ b1. (We remark that 0 ∈ N.) This is a total ordering, i.e. for any<br />

pair z1, z2 ∈ N × N, either z ≤ z1 or z1 ≤ z. For any finite tuple<br />

a1, a2, . . .,an<br />

of elements from Z, let the degree of a1, a2, . . .,an be the element (a, l) ∈ N × N<br />

defined as follows: a = max {|a1|, |a2|, . . ., |an|} and l is the number of times a<br />

occurs in the list |a1|, |a2|, . . ., |an|. The proof consists in establishing the following<br />

assertion :<br />

A) When (a, l) ∈ N × N and the conclusion of the lemma holds for all finite<br />

tuples from Z whose degree is strictly less than (a, l), then the conclusion<br />

of the lemma holds for all finite tuples from Z whose degree is (a, l).<br />

If this assertion has been proved the lemma follows by a simple contradiction<br />

argument as follows : If there was a finite tuple a1, a2, . . .,an of elements from Z<br />

for which the lemma failed, then we could use A) repeatedly to conclude that there<br />

would be such a tuple, for which the lemma failed, with degree = (0, 1). But the<br />

lemma is obviously true for such tuples, and we have the desired contradiction.<br />

In order to prove A) we first establish the lemma in the special case where ai ≥ 0<br />

for all i. In this case we have that<br />

0 ≤ xi ≤ aixi ≤ a1x1 + · · · + anxn = 0<br />

when ai > 0, so we see that ai > 0 ⇒ xi = 0 in this case. If xi = 0 for all<br />

i we can set m = 1, y1 = 0 and bi1 = 0 for all i = 1, 2, . . ., n, i.e. the lemma<br />

holds in this case. We can therefore assume that ai = 0 for at least one i. After<br />

reindexing there is then an m ∈ {1, 2, . . ., n} such that a1 = a2 = · · · = am = 0<br />

and xj = 0, j = m + 1, m + 2, . . .,n. By choosing yj = xj, j = 1, 2, . . ., m, and<br />

bjj = 1, j = 1, 2, . . ., m, and bij = 0 for all other i, j, we get the conclusion of the<br />

lemma in this case also.<br />

The case where ai ≤ 0 for all i follows from the case where ai ≥ 0 for all i by<br />

multiplying the ai’s by −1.<br />

Now we proceed to the proof of A), i.e. we consider a general finite tuple<br />

a1, a2, . . ., an and assume that the conclusion of the lemma holds for all tuples whose<br />

degree is strictly less than that of the given one. If the degree is (0, l) for some l,<br />

we are in the case where ai = 0 for all i and this case has already been covered.<br />

So we may assume that the degree is (a, l) with a > 0. Furthermore, by what we<br />

have just shown, we can disgard the case where all the ai’s are either non-negative<br />

or non-positive, i.e. we may assume that ai > 0 for some i and aj < 0 for some j.<br />

Reindex the ai’s such that |a1| = a. As it is harmless to multiply by −1, we<br />

can assume that a1 = a. Next reindex a2, a3, . . .,an and x2, x3, . . .,xn so that<br />

a1, a2, . . ., ar ≥ 0 and ar+1, ar+2, . . .,an < 0 for some r ∈ {2, 3, . . ., n−1}. Note that<br />

a ≥ −ai ≥ 0 for all i = r + 1, r + 2, . . ., n. We have<br />

ax1 = a1x1 ≤ a1x1 + a2x2 + · · · + arxr =<br />

− ar+1xr+1 − ar+2xr+2 − · · · − anxn ≤ a(xr+1 + xr+2 + · · · + xn),


which implies that<br />

3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 87<br />

x1 ≤ xr+1 + xr+2 + · · · + xn<br />

because G is unperforated. By n−r applications of the Riesz decomposition property<br />

we can write<br />

x1 = zr+1 + zr+2 + · · · + zn,<br />

where 0 ≤ zi ≤ xi, i = r + 1, r + 2, . . .,n. Then<br />

(a1 + ar+1)zr+1 + · · · + (a1 + an)zn + a2x2 + · · · + arxr+<br />

ar+1(xr+1 − zr+1) + · · · + an(xn − zn) =<br />

a1(zr+1 + · · · + zn) + a2x2 + · · · + anxn = 0.<br />

As 0 ≤ a1 + ai < a1 = a for all i = r + 1, r + 2, . . .,n, we see that a appears in<br />

the list<br />

|a1 + ar+1|, . . ., |a1 + an|, |a2|, . . .,|an|<br />

at most l − 1 times. Consequently the degree of the tuple<br />

a1 + ar+1, . . .,a1 + an, a2, a3, . . .,an<br />

is strictly less than (a, l). Hence, by assumption, we can find elements y1, y2, . . .,ym ∈<br />

G + and nonnegative integers cij, i = r + 1, r + 2, . . .,n, j = 1, 2, . . ., m, and dij, i =<br />

2, 3, . . ., n, j = 1, 2, . . ., m, such that<br />

zi = ci1y1 + · · · + cimym, i = r + 1, . . ., n,<br />

while<br />

xi = di1y1 + · · · + dimym, i = 2, 3, . . ., r,<br />

xi − zi = di1y1 + · · · + dimym, i = r + 1, . . ., n.<br />

(a1 + ar+1)cr+1,j + · · · + (a1 + an)cnj + a2d2j + · · · + andnj = 0<br />

for all j = 1, 2, . . .,m.<br />

Define bij by<br />

i=1<br />

b1j = cr+1,j + · · · + cnj, j = 1, 2, . . ., m,<br />

bij = dij, i = 2, 3, . . ., r, j = 1, 2, . . ., m,<br />

bij = cij + dij, i = r + 1, . . ., n, j = 1, . . ., m.<br />

Then xi = bi1y1 + · · · + bimym for all i = 1, 2, . . ., n, and<br />

n <br />

n <br />

r n<br />

n<br />

aibij = a1 cij + aidij+ ai(cij+dij) =<br />

i=r+1<br />

i=2<br />

i=r+1<br />

i=r+1<br />

(a1+ai)cij+<br />

n<br />

aidij = 0<br />

for all j = 1, 2, . . ., m. This completes the proof of A) and hence of the lemma. <br />

Lemma 3.42. Let G be a dimension group, H1 a simplicial group, and let h1 :<br />

H1 → G be a positive homomorphism. Then there is a simplicial group H2 and<br />

positive homomorphisms h : H1 → H2 and h2 : H2 → G such that h1 = h2 ◦ h and<br />

ker h1 = ker h.<br />

Proof. We will first prove the following assertion by induction in k :<br />

(A) Let H be a simplicial group and h : H → G a positive homomorphism.<br />

Let v1, v2, · · · , vk be elements of ker h. There is then a simplicial group H ′ and<br />

positive homomorphisms, h ′ : H → H ′ and h ′′ : H ′ → G such that h = h ′′ ◦ h ′ and<br />

{v1, v2, · · · , vk} ⊆ ker h ′ .<br />

i=2


88 1. FUNDAMENTALS<br />

We start the induction by considering the case k = 1 : If H = {0} we can just<br />

take H ′ = H, h ′′ = h ′ = 0. Now assume that H = {0}. Let e1, e2, . . .,en be elements<br />

of H such that<br />

H = Ze1 ⊕ Ze2 ⊕ · · · ⊕ Zen<br />

and<br />

H + = Ne1 + Ne2 + · · · + Nen.<br />

Such elements exist because H is a simplicial group. Set xi = h(ei) ∈ G + , i =<br />

1, 2, . . ., n. Write v1 = a11e1 + · · · + a1nen for some a1i ∈ Z. Then<br />

a11x1 + · · · + a1nxn = h(a11e1 + · · · + a1nen) = h(v1) = 0.<br />

By Lemma 3.41 there exist elements y1, y2, . . ., ym ∈ G + and non-negative integers<br />

bij, i = 1, 2, . . ., n, j = 1, 2, . . ., m, such that<br />

and<br />

xi = bi1y1 + · · · + bimym, i = 1, 2, . . ., n,<br />

a11b1j + · · · + a1nbnj = 0, j = 1, 2, . . .,m.<br />

Let H ′ be the simplicial group H ′ = Z m , and let fi, i = 1, 2, . . ., m, be the standard<br />

basis for Z m . We may define homomorphisms h ′ : H → H ′ and h ′′ : H ′ → G so that<br />

and<br />

h ′ (ei) = bi1f1 + · · · + bimfm, i = 1, 2, . . .,n,<br />

h ′′ (fj) = yj, j = 1, 2, . . ., m.<br />

As each bij ≥ 0 and each yj ≥ 0, we see that both h ′ and h ′′ are positive homomor-<br />

phisms. Since<br />

h ′′ ◦ h ′ (ei) = h ′′ (<br />

m<br />

bijfj) =<br />

j=1<br />

for all i, we see that h ′′ ◦ h ′ = h. In addition,<br />

h ′ (v1) =<br />

n<br />

i=1<br />

a1ih ′ (ei) =<br />

n<br />

i=1<br />

m<br />

j=1<br />

bijyj = xi<br />

m<br />

a1ibijfj = 0,<br />

so that v1 ∈ ker h ′ as desired. Thus (A) holds when k = 1.<br />

Now let k > 1 and assume that there exists a simplicial group H ′ and positive<br />

homomorphisms h ′ : H → H ′ and h ′′ : H ′ → G such that h = h ′′ ◦ h ′ and<br />

{v1, v2, . . .,vk−1} ⊆ ker h ′ . Note that h ′′ ◦h ′ (vk) = h(vk) = 0 so that h ′ (vk) ∈ ker h ′′ .<br />

Applying the case k = 1, we obtain a simplicial group H ′′ and positive homomorphisms<br />

ψ : H ′ → H ′′ and l : H ′′ → G such that h ′′ = l ◦ψ and h ′ (vk) ∈ ker ψ. Then<br />

ψ ′ = ψ ◦ h ′ : H → H ′′ is a positive homomorphism such that l ◦ ψ ′ = h ′′ ◦ h ′ = h.<br />

Clearly, {v1, . . .,vk−1} ⊆ ker ψ ′ and ψ ′ (vk) = ψ ◦ h ′ (vk) = 0. This completes the<br />

induction step.<br />

Using assertion (A) we can easily prove the lemma. Since H1 is a finitely generated<br />

Abelian group, so is ker h1. Thus<br />

j=1<br />

ker h1 = Zv1 + Zv2 + · · · + Zvk<br />

for some elements v1, v2, . . .,vk ∈ ker h1. There is then a simplicial group H2 and<br />

positive homomorphisms h : H1 → H2 and h2 : H2 → G such that h1 = h2 ◦ h<br />

and {v1, v2, · · · , vk} ⊆ ker h. It is then automatic that ker h1 ⊆ ker h and the<br />

factorization h1 = h2 ◦ h shows that ker h ⊆ ker h1.


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 89<br />

Theorem 3.43. Let G be a countable dimension group. There is then a sequence<br />

G1<br />

p1 <br />

G2<br />

p2 <br />

G3<br />

p3 <br />

. . .<br />

of simplicial groups and positive homomorphisms such that G ≃ lim<br />

−→ (Gn, pn) as partially<br />

ordered Abelian groups.<br />

Proof. Let G + = {x1, x2, . . . } be a numeration of G + . We shall construct<br />

simplicial groups G1, G2, . . . and positive homomorphisms gn : Gn → G and pn :<br />

Gn → Gn+1 for all n such that<br />

1) xn ∈ gn(G + n) for all n,<br />

2) gn ◦ pn−1 = gn−1 for all n > 1,<br />

3) ker gn = ker pn for all n.<br />

To start, set G1 = Z and define g1 : G1 → G such that g1(1) = x1.<br />

Now assume that G1, G2, . . .,Gn, g1, g2, . . ., gn, p1, p2, . . .,pn−1 have been constructed.<br />

Set H = Gn ⊕ Z, ordered by G + n ⊕ N, so that H is a simplicial group. We<br />

may define a positive homomorphism h : H → G by<br />

h(x, z) = gn(x) + zxn+1.<br />

Applying Lemma 3.42, we obtain a simplicial group Gn+1 and positive homomorphisms<br />

h ′ : H → Gn+1 and gn+1 : Gn+1 → G such that h = gn+1 ◦ h ′ and<br />

ker h = ker h ′ . As (0, 1) ∈ H + , we have h ′ ((0, 1)) ∈ G + n+1 with gn+1 ◦ h ′ ((0, 1)) =<br />

h((0, 1)) = xn+1, so that xn+1 ∈ gn+1(G + n+1 ). Let q : Gn → H be the natural<br />

injection (to the first coordinate), and set pn = h ′ ◦ q. Then pn is a positive homomorphism<br />

Gn → Gn+1 such that gn+1 ◦ pn = gn+1 ◦ h ′ ◦ q = h ◦ q = gn. In<br />

particular, ker pn ⊆ ker gn. On the other hand, q(ker gn) ⊆ ker h = ker h ′ , and<br />

hence ker gn ⊆ ker h ′ ◦ q = ker pn. Thus ker gn = ker pn as desired, and we have<br />

completed the induction step.<br />

Now let K = lim(Gn,<br />

pn) be the partially ordered inductive limit group and let<br />

−→<br />

kn : Gn → K be the natural map. Because of condition 2), there is a unique positive<br />

homomorphism g : K → G such that g ◦ kn = gn for all n. Condition 1) then shows<br />

that xn ∈ gn(G + n ) = g ◦ kn(G + n ) ⊆ g(K+ ) for all n, i.e. G + = g(K + ). In particular,<br />

it follows that g is surjective. To show that g is also injective, let x ∈ ker g. Write<br />

x = kn(y) for some n ∈ N and some y ∈ Gn. Then gn(y) = g ◦ kn(y) = 0. Condition<br />

3) now implies that y ∈ ker pn so that x = kn+1 ◦ pn(y) = 0. <br />

Let G be a partially ordered Abelian group. An element u ∈ G + is an order unit<br />

when the following holds : For all g ∈ G there is an n ∈ N such that g ≤ nu. Thus<br />

the element [1] in the K0-group of a unital C ∗ -algebra is always an order unit and<br />

it is through this example the notion of an order unit becomes important here.<br />

Proposition 3.44. Let G be a countable dimension group and u ∈ G an order<br />

unit. There is then a sequence<br />

G1<br />

p1 <br />

G2<br />

p2 <br />

G3<br />

p3 <br />

. . .<br />

of simplicial groups with order units un ∈ Gn, positive homomorphisms pn such that<br />

pn(un) = un+1, and an isomorphism ϕ : lim(Gn,<br />

pn) → G of partially ordered Abelian<br />

−→<br />

groups such that<br />

ϕ(qn(un)) = u, n ∈ N,<br />

when qn : Gn → lim(Gn,<br />

pn) is the canonical map.<br />

−→


90 1. FUNDAMENTALS<br />

Proof. We know from Theorem 3.43 that there is a sequence<br />

H1<br />

r1 <br />

H2<br />

r2 <br />

H3<br />

r3 <br />

. . .<br />

of simplicial groups and positive homomorphisms such that G ≃ lim(Hn,<br />

rn) as par-<br />

−→<br />

tially ordered groups. Let v ∈ lim(Hn,<br />

rn) be the image of u under this isomorphism.<br />

−→<br />

Then v is an order-unit in lim(Hn,<br />

rn). In particular, v is positive, i.e. v ∈ q<br />

−→ ′ n(H + n )<br />

for some n ∈ N. Here q ′ n : Hn → lim(Hn,<br />

rn) is the canonical map. Since we can<br />

−→<br />

ignore Hi, i < n, we may assume that v = q ′ 1 (u1) for some u1 ∈ H + 1 . Define un ∈ Hn<br />

by<br />

Set<br />

un+1 = rn(un), n ∈ N.<br />

Gn = {x ∈ Hn : ∃k ∈ N : −kun ≤ x ≤ kun},<br />

n ∈ N. It is straightforward to check that Gn is a simplicial group in the ordering<br />

inherited from Hn, i.e. when the positive semigroup is G + n = H + n ∩ Gn, and that un<br />

is an order unit for Gn. Set pn = rn|Gn. Then<br />

G1<br />

p1 <br />

G2<br />

p2 <br />

G3<br />

p3 <br />

. . .<br />

is a sequence of simplicial groups with positive connecting maps such that pn(un) =<br />

un+1 for all n. Let qn : Gn → lim(Gn,<br />

pn) be the canonical maps and note that there<br />

−→<br />

is a positive group homomorphism<br />

ψ : lim(Gn,<br />

pn) → lim(Hn,<br />

rn)<br />

−→ −→<br />

such that ψ ◦ qn = q ′ n |Gn. It is clear that ψ is injective and that ψ(qn(un)) = v for<br />

all n. It now suffices to show that ψ lim(Gn,<br />

pn)<br />

−→ + = lim(Hn,<br />

rn)<br />

−→ + . So let x = q ′ n (y)<br />

for some y ∈ H + n<br />

. Since v is an order unit in lim<br />

−→ (Hn, rn), there is a k ∈ N such that<br />

0 ≤ x ≤ kv, i.e. 0 ≤ q ′ n (y) ≤ kq′ 1 (u1) = kq ′ n (un). Since q ′ n (kun − y) ≥ 0, there is<br />

an m ≥ n such that rm,n(kun − y) ≥ 0. Then −kum ≤ 0 ≤ rm,n(y) ≤ kum, so that<br />

rm,n(y) ∈ H + m ∩ Gm = G + m. Hence qm(rm,n(y)) ∈ lim<br />

−→ (Gn, pn) + and ψ(qm(rm,n(y))) =<br />

q ′ m(rm,n(y)) = q ′ n(y) = x.<br />

<br />

Theorem 3.45. Let G be a countable dimension group and u ∈ G an order unit.<br />

There is then a unital AF-algebra A and an isomorphism ϕ : K0(A) → G of partially<br />

ordered groups such that ϕ([1]) = u.<br />

In short,<br />

(G, u) = (K0(A), [1]).<br />

Proof. By Proposition 3.44 we may assume that G is the inductive limit of a<br />

sequence<br />

G1<br />

p1 <br />

G2<br />

p2 <br />

G3<br />

p3 <br />

. . .<br />

of simplicial groups with order units un ∈ Gn such that pn(un) = un+1 and qn(un) =<br />

u for all n. In fact, we may assume that there is a sequence {Nn} in N such that<br />

) and each<br />

Gn = ZNn , with its simplicial ordering. For each n, un = (dn 1 , dn2 , . . .,dn Nn<br />

dn i > 0 because un is an order unit in ZNn . Each pn : ZNn → ZNn+1 is given by a


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 91<br />

Nn+1 ×Nn matrix {sij} with entries from N. Note that the condition pn(un) = un+1<br />

yields that<br />

Nn <br />

Set<br />

j=1<br />

sijd n j<br />

= dn+1<br />

i , i = 1, 2, . . ., Nn+1.<br />

An = Md n 1 ⊕ Md n 2 ⊕ · · · ⊕ Md n Nn<br />

and note that because of (2.8) and (2.9) there is a unital standard homomorphism<br />

ϕn : An → An+1 for each n. Set A = lim<br />

(An, ϕn). By combining Exercise 3.50 with<br />

−→<br />

Theorem 3.21 one sees that there is an isomorphism ϕ : K0(A) → G = lim(Gn,<br />

pn)<br />

−→<br />

such that ϕ([1]) = u. <br />

We now turn to the question of which subsets Σ of a dimension group G can<br />

be the scale of an AF-algebra A with G ≃ K0(A). When we restrict the attention<br />

to unital AF-<strong>algebras</strong>, Lemma 3.44 and Theorem 3.45 tells us the full story; the<br />

possible subsets Σ are exactly those of form Σ = {g ∈ G : 0 ≤ g ≤ e} for some order<br />

unit e in G. But what about the non-unital case?<br />

Definition 3.46. Let G be a dimension group. A subset Σ ⊆ G + is called a<br />

scale when the following hold :<br />

1) {g1 + g2 + · · · + gn : n ∈ N, g1, g2, · · · , gn ∈ Σ} = G + ,<br />

2) 0 ≤ h ≤ g ∈ Σ ⇒ h ∈ Σ, g, h ∈ G,<br />

3) For every pair g, h ∈ Σ there is an element k ∈ Σ such that g ≤ k and<br />

h ≤ k.<br />

Theorem 3.47. Let Σ be a subset of a countable dimension group G. There<br />

is then an AF-algebra A and an isomorphism f : K0(A) → G of partially ordered<br />

groups such that f(Σ(A)) = Σ if and only if Σ is a scale.<br />

Proof. To prove the ’only if’ part we must show that Σ(A) is a scale in K0(A)<br />

when A is an AF algebra. We check that conditions 1),2) and 3) of Definition 3.46<br />

hold. Let A be the inductive limit of the sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

of finite dimensional C ∗ -<strong>algebras</strong> An. Let x ∈ K0(A) + . By Theorem 3.21 there is an<br />

N ∈ N and a y ∈ K0(AN) + such that x = ϕ∞,N ∗ (y), where ϕ∞,N : AN → A is the<br />

canonical map. Since AN is finite dimensional, it follows from Exercise 3.50 that y<br />

is in the semi-group of K0(AN) generated by Σ(AN). Since ϕ∞,N ∗ (Σ(AN)) ⊆ Σ(A),<br />

we see that x is in the semi-group of K0(A) generated by Σ(A). In particular, 1)<br />

holds. Let next x, y ∈ K0(A) such that 0 ≤ y ≤ x ∈ Σ(A). To check that 2)<br />

holds, note first that x = [p] for some projection p ∈ A. By Lemma 3.15 there is<br />

. Thus p and<br />

an N ∈ N and a projection q ∈ AN such that µ∞,N(q) − p < 1<br />

3<br />

ϕ∞,N(q) are equivalent by Lemma 3.14, i.e. x = [p] = [ϕ∞,N(q)]. Since 0 ≤ y, there<br />

is, by Theorem 3.21, a z ∈ K0(AM) + for some M ≥ N such that y = ϕ∞,M ∗ (z).<br />

Since y ≤ x = [ϕ∞,N(q)], there is a L ≥ M such that 0 ≤ ϕL,M ∗ (z) ≤ ϕL,N ∗ ([q])] in<br />

K0(AL). Since [ϕL,N(q)] ∈ Σ(AL), we see from Exercise 3.50 that ϕL,M ∗ (z) ∈ Σ(AL).<br />

Hence y = ϕ∞,L∗ (ϕL,M ∗ (z)) ∈ Σ(A). Thus 2) holds. To check 3), let x, y ∈ Σ(A).<br />

As above we see from Lemma 3.14 and Lemma 3.15 that there is an N ∈ N and<br />

projections p, q ∈ AN such that x = ϕ∞,N ∗ ([p]), y = ϕ∞,N ∗ ([q]). Let e be the unit


92 1. FUNDAMENTALS<br />

of AN. Then p ≤ e and q ≤ e in AN and hence z = ϕ∞,N ∗ ([e]) ∈ Σ(A) satisfies that<br />

x ≤ z, y ≤ z. Thus 3) holds also.<br />

To prove the ’if’ part, assume that Σ is a scale. Let e1, e2, e3, . . . be a numbering<br />

of the elements in Σ. It follows from 3) of Definition 3.46 that there is a sequence<br />

f1 ≤ f2 ≤ f3 ≤ . . . of elements in Σ such that fn ≥ ej, j = 1, 2, . . ., n, for all n ∈ N.<br />

Set<br />

Hn = {g ∈ G : −kfn ≤ g ≤ kfn for some k ∈ N}.<br />

It is straightforward to check that Hn is a dimension group for which fn is an<br />

order unit. By Theorem 3.45 there is a unital AF-algebra An and an isomorphism<br />

ψn : K0(An) → Hn of partially ordered groups such that ψn([1]) = fn. Then<br />

ψ −1<br />

n+1 ◦ ψn : K0(An) → K0(An+1) is a positive group homomorphism taking Σ(An)<br />

into Σ(An+1). By Theorem 3.35 1) there is a ∗-homomorphism ϕn : An → An+1<br />

such that ϕn∗ = ψ −1<br />

n+1 ◦ ψn. Let A be the inductive limit C ∗ -algebra of the sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

. By Corollary 2.17 A is an AF-algebra. Since<br />

Hn<br />

<br />

⏐<br />

⏐ψn<br />

K0(An) −−−→<br />

ϕn∗<br />

ϕ3 <br />

· · ·<br />

⊆<br />

−−−→ Hn+1<br />

<br />

⏐<br />

⏐ψn+1<br />

K0(An+1)<br />

commutes, we get a group homomorphism ψ : K0(A) = lim<br />

−→ (K0(An), ϕn∗) → G such<br />

that ψ ◦ ϕ∞,n ∗ = ψn for all n. Since each ψn is injective, so is ψ. From the first part<br />

of the proof we know that an element x ∈ Σ(A) is of the form x = ϕ∞,N ∗ (y) for<br />

some N ∈ N and some y ∈ Σ(AN). Since ψ(x) = ψN(y) ∈ {h ∈ G : 0 ≤ h ≤ fN},<br />

it follows from condition 2) of Definition 33.46 that ψ(x) ∈ Σ. This shows that<br />

ψ(Σ(A)) ⊆ Σ. On the other hand, if g ∈ Σ then 0 ≤ g ≤ fn for some n and hence<br />

g ∈ Hn. Since ψn(Σ(An)) = {h ∈ Hn : 0 ≤ h ≤ fn} there is a y ∈ Σ(An) such that<br />

ψn(y) = g. Then x = ϕ∞,n ∗ (y) ∈ Σ(A) and ψ(x) = ψn(y) = g. Hence ψ(Σ(A)) = Σ.<br />

Since Σ generates G by condition 1) of Definition 3.46, this implies in particular<br />

that ψ is surjective. <br />

3.4. Exercises to Chapter 3.<br />

Exercise 3.48. Let Tr : Mn → C be the usual trace.<br />

1) Show that there is a linear map Trk : Mk(Mn) → C given by<br />

⎛<br />

a11<br />

⎜a21<br />

Trk<br />

⎜<br />

⎝<br />

.<br />

a12<br />

a22<br />

.<br />

⎞<br />

. . . a1k<br />

. . . a2k⎟<br />

.<br />

⎟<br />

.. .<br />

⎠ =<br />

k<br />

Tr(aii)<br />

ak1 ak2 . . . akk<br />

which satisfies that Trk(AB) = Trk(BA), A, B ∈ Mk(Mn).<br />

2) Show that there is a semi-group isomorphism T : V (Mn) → N such that<br />

T([e]) = Trk(e), e ∈ Pk(Mn).<br />

3) Show that K00(Mn) ≃ Z.<br />

4) Show that under this isomorphism K00(Mn) + = N.<br />

i=1


3. K0 AND THE CLASSIFICATION OF AF-ALGEBRAS 93<br />

Exercise 3.49. Let A = lim<br />

−→ (An, ϕn) be the inductive limit of the sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

. . .<br />

of unital C ∗ -<strong>algebras</strong>, but not necessarily with unital connecting maps. Show that<br />

ιA∗ : K00(A) → K0(A) is an isomorphism of pre-ordered Abelian groups. (Hint :<br />

Combine Theorem 3.21, Theorem 3.16 and Lemma 3.19)<br />

Exercise 3.50. Let A = Mn1⊕Mn2⊕· · ·⊕MnN and let {ed ij : i, j = 1, 2, . . ., nd, d =<br />

1, 2, . . ., N} be the standard matrix units in A. Show that the map Z N → K0(A)<br />

given by<br />

is an isomorphism. Show also that<br />

and<br />

Σ(A) = {<br />

(z1, z2, . . ., zN) ↦→<br />

N<br />

i=1<br />

zi[e i 11 ]<br />

K0(A) + = N[e 1 11] + N[e 2 11] + · · · + N[e N 11]<br />

N<br />

i=1<br />

ti[e i 11 ] : ti ∈ {0, 1, 2 . . ., ni}, i = 1, 2, . . ., N}. <br />

Exercise 3.51. Let G be a partially ordered Abelian group. Assume that G<br />

has the Riesz interpolation property. Show that G also has the following property:<br />

When x1, x2, · · · , xn and y1, y2, · · · , ym are elements of G such that xi ≤ yj for all<br />

i, j, there is an element z ∈ G such that xi ≤ z ≤ yj for all i, j.<br />

Exercise 3.52. Determine the Grothendieck group G(S) for the following semigroups<br />

:<br />

(1) S = N with +.<br />

(2) S = {n ∈ N : n > 7} with +.<br />

(3) S = [−1, 1] with · (multiplication).<br />

(4) S = [−1, 1]\{0} with ·.<br />

(5) S = {z ∈ C : Rez ≥ 0, Imz ≤ 0} with +.<br />

Exercise 3.53. a) Any subgroup of R is a partially ordered Abelian group with<br />

the order inherited from R.<br />

b) In G = Z ⊕ Z the following subsets all make G into a pre-ordered group :<br />

1) G + = {(x, y) ∈ Z 2 : x ≥ 0, y ≥ 0}.<br />

2) G + = {(x, y) ∈ Z 2 : x ≥ 0}.<br />

3) G + = {(x, y) ∈ Z 2 : x > 0 or x = 0, y ≥ 0}.<br />

In the second case the pre-order is not a partial order.<br />

c) Construct 7 examples of pre-ordered groups yourself.<br />

Exercise 3.54. Let A be a separable C ∗ -algebra. Prove that K00(A) and K0(A)<br />

are countable groups. (Hint : Remember that projections that are close are equivalent.)<br />

Exercise 3.55. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·


94 1. FUNDAMENTALS<br />

be a sequence of C ∗ -<strong>algebras</strong> and A = lim<br />

−→ (An, ϕn). For each d ∈ N we have a new<br />

sequence of C ∗ -<strong>algebras</strong>,<br />

Md(A1) ϕ1 <br />

Md(A2) ϕ2 <br />

Md(A3) ϕ3 <br />

· · ·<br />

. Prove that lim<br />

−→ (Md(An), ϕn) ≃ Md(A).<br />

Exercise 3.56. Let ϕ, ψ : A → B be ∗-homomorphisms between C ∗ -<strong>algebras</strong>.<br />

Assume that there is a sequence {un} of unitaries in B + such that ϕ(a) = limn→∞ unψ(a)u ∗ n<br />

for all a ∈ A. Prove that ϕ∗ = ψ∗ : K0(A) → K0(B).<br />

Exercise 3.57. Let P be the projection introduced in Example 3.22. Show that<br />

[P] − [( 0 1 )] is not zero K00(A).<br />

Exercise 3.58. Let A be a <strong>C∗</strong>-algebra and define s : A → Mn(A) by<br />

⎛ ⎞<br />

a 0 . . . 0<br />

⎜0<br />

0 . . . 0⎟<br />

s(a) = ⎜<br />

⎝ .<br />

⎟<br />

. . .. 0<br />

⎠<br />

0 0 . . . 0<br />

.<br />

Show that s∗ : K00(A) → K00(Mn(A)) is an isomorphism.<br />

Exercise 3.59. Let A, B be C ∗ -<strong>algebras</strong> and let p1 : A⊕B → A and p2 : A⊕B →<br />

B be the two projections. Show that the map K0(A ⊕ B) ∋ x → (p1∗(x), p2∗(x)) ∈<br />

K0(A) ⊕ K0(B) gives an isomorphism of pre-ordered groups when<br />

(K0(A) ⊕ K0(B)) + = K0(A) + ⊕ K0(B) + .<br />

Exercise 3.60. a) Let A and B be C ∗ -<strong>algebras</strong> and ϕt : A → B, t ∈ [0, 1], be a<br />

family of ∗-homomorphisms such that t ↦→ ϕt(a) is continuous for all a ∈ A. Prove<br />

that ϕ0∗ = ϕ1∗ : K00(A) → K00(B) and ϕ0∗ = ϕ1∗ : K0(A) → K0(B).<br />

b) Define k : C → C[0, 1] by k(λ)(s) = λ. Prove that k∗ : K0(C) → K0(C[0, 1])<br />

is an isomorphism of partially ordered groups. (Hint : Define ψ : C[0, 1] → C<br />

by ψ(f) = f(0). Then ψ ◦ k = idC. Conclude that k∗ is injective. Define<br />

ϕt : C[0, 1] → C[0, 1] by ϕt(f)(s) = f(st), s ∈ [0, 1] and use this to show that k∗<br />

is surjective.)<br />

c) Show that K0(C([0, 1], A)) ≃ K0(A) as pre-ordered Abelian groups.<br />

Exercise 3.61. Determine the ordered K0-group of a UHF-algebra.<br />

Exercise 3.62. Determine the ordered K0-group of C(X), where X is the Cantor<br />

set, cf. Exercise 2.24<br />

4. Choquet simplices and C ∗ -<strong>algebras</strong><br />

4.1. Compact convex sets and functional analysis. A compact convex set<br />

K is a convex subseteq of a locally convex real vector space which is compact in the<br />

relative topology. In this section we will develop the basic connection between the<br />

theory of compact convex sets and functional analysis.<br />

If A is a C ∗ -algebra, it follows from the Banach-Alouglu theorem that the unit<br />

ball of A ∗ is compact in the weak ∗ topology. If A has a unit, it follows that<br />

S(A) = {ω ∈ A ∗ : ω = ω(1) = 1}


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 95<br />

is a compact convex subset of A ∗ . By Lemma 1.79 S(A) consists exactly of the<br />

states of A, i.e. of the positive linear functionals of norm 1 on A. Consequently,<br />

S(A) is called the state space of A. By the Krein-Milman theorem every compact<br />

convex set is the closed convex hull of its extreme points, so this is in particular<br />

the case for S(A). The extreme points of S(A) are called pure states of A. It is<br />

symptomatic for the close relationship between the theory of compact convex sets<br />

and the theory of C ∗ -<strong>algebras</strong> that the former commences by determining the pure<br />

states of C(X) for a compact Hausdorff space X.<br />

Theorem 4.1. Let X be a compact Hausdorff space. A pure state of C(X) is of<br />

the form C(X) ∋ f ↦→ f(x) for some point x in X.<br />

Proof. Let ω be a pure state of C(X). We first prove that ω is a character,<br />

i.e. that ω(fg) = ω(f)ω(g), f, g ∈ C(X). It clearly suffices to show this for<br />

non-negative f and g with f ≤ 1. Set µ(h) = ω(fh), h ∈ C(X). Since f ≥ 0,<br />

µ is a positive linear functional. If µ = 0, we have in particular that µ(g) =<br />

ω(fg) = 0 = ω(g)µ(1) = ω(g)ω(f). If µ = ω, we have that ω(f) = 1 and that<br />

ω(fg) = µ(g) = ω(g) = ω(f)ω(g). If µ = 0 and µ = ω, we have that<br />

ω = µ(1) µ<br />

ω − µ<br />

+ (ω(1) − µ(1))<br />

µ(1) ω(1) − µ(1) .<br />

Note that µ<br />

µ(1) ,<br />

ω−µ<br />

ω(1)−µ(1)<br />

∈ S(C(X)). Since µ(1), ω(1) −µ(1) ∈ [0, 1], µ(1)+(ω(1) −<br />

µ(1)) = 1, the assumption that ω is pure, i.e. is an extreme point in S(C(X)), implies<br />

that ω = µ<br />

µ(1)<br />

. In particular these states must agree on g, i.e. ω(g) = µ(g)<br />

µ(1)<br />

= ω(fg)<br />

ω(f) .<br />

Since ω is a character, ker ω is a twosided ideal Iω in C(X). Note that C(X)/Iω ≃<br />

C. Set Y = {y ∈ X : f(y) = 0, f ∈ Iω}. We assert that Y is not empty. Indeed,<br />

if it was we could for any z ∈ X find a function gz ∈ Iω such that gz(z) = 0. Then<br />

|gz|(y) > 0 for all y in an open neighbourhood Vz of z and |gz| = (gzgz) 1<br />

2 ∈ Iω. By<br />

compactness of X this would give us a finite set gz1, gz2, . . .,gzn of such functions<br />

such that n<br />

i=1 |gzi |(y) > 0 for all y ∈ X. Then h = n<br />

i=1 |gzi | ∈ Iω and hh −1 = 1.<br />

Thus 0 = ω(h)ω(h −1 ) = ω(1) = 1, contradiction. Hence Y = ∅ as asserted. Let x<br />

be an arbitrary point in Y . Then, for any f ∈ C(X), f − f(x)1 ∈ Iω, so ω(f) =<br />

ω(f − f(x)1) + ω(f(x)1) = f(x). <br />

Exercise 4.2. Prove the converse of Theorem 4.1 : For every x ∈ X the functional<br />

f ↦→ f(x) is a pure state of C(X).<br />

We have now, in effect, proved that the category of compact Hausdorff spaces is<br />

equivalent to the category of unital abelian C ∗ -<strong>algebras</strong> : To a compact Hausdorff<br />

space X we can associate the unital abelian C ∗ -algebra C(X), cf. Example 1.20.<br />

On the other hand, we can to any unital abelian C ∗ -algebra associate the compact<br />

Hausdorff space of characters of the algebra. By Theorem 1.31 and Theorem 4.1,<br />

these two operations are inverses of each other.<br />

In the right perspective the connection between the theory of compact convex sets<br />

and functional analysis is a generalisation of the bijective correspondance of compact<br />

Hausdorff spaces with unital abelian C ∗ -<strong>algebras</strong>. When K is a compact convex set,<br />

Aff K will denote the continuous realvalued affine functions on K, i.e. Aff K consists<br />

of the continuous functions f : K → R which respect the convex structure of K in<br />

the sense that f(tx + (1 − t)y) = tf(x) + (1 − t)f(y), t ∈ [0, 1], x, y ∈ K. Aff K<br />

is clearly a closed subspace of the continuous realvalued functions CR(K), i.e. of


96 1. FUNDAMENTALS<br />

the selfadjoint part of the C ∗ -algebra C(K). The state space , S(Aff K), of this<br />

subspace, i.e.<br />

S(Aff K) = {ω ∈ Aff K ∗ : ω = ω(1) = 1}<br />

is a compact convex subset of Aff K ∗ in the weak ∗ topology by the Banach-Alouglu<br />

theorem. The following theorem shows that S(Aff K) is K again, so that the real<br />

Banach space Aff K contains all information about K. Note that every point x ∈ K<br />

defines an element ωx of S(Aff K) by<br />

ωx(f) = f(x), f ∈ Aff K.<br />

The set of extreme points in a compact convex set K will be denoted by ∂eK.<br />

Theorem 4.3. The map x ↦→ ωx is an affine homeomorphism from K onto<br />

S(Aff K).<br />

Proof. Set Φ(x) = ωx. If {xα} is a net in K converging to x ∈ K, then<br />

limα f(xα) = f(x) for all f ∈ CR(K), in particular for all f ∈ Aff K. This shows<br />

that Φ is continuous. Assume x, y ∈ K, x = y. Let E be the locally convex<br />

real vectorspace of which K is a compact convex subseteq. By the Hahn-Banach<br />

separation theorem there is a continuous linear functional ω on E such that ω(x −<br />

y) = 0. The restriction ω|K defines an element of Aff K which takes different values<br />

at x and y. Thus ωx = ωy, showing that Φ is injective.<br />

To prove that Φ is also surjective, let first ω be an extreme point in S(Aff K).<br />

By the Hahn-Banach extension theorem there is a continuous real functional ω1 on<br />

CR(K) with the same norm as ω such that ω1|Aff K = ω. Extend ω1 to a linear<br />

functional ω2 on C(K) by<br />

ω2(f) = ω1( 1<br />

1<br />

(f + f)) + iω1( (f − f)), f ∈ C(K).<br />

2 2i<br />

Clearly, ω2 ≤ 2ω1 = 2ω. If 0 ≤ f ≤ 1, then ω2(1 − f) ≤ |ω2(1 − f)| ≤<br />

ω11 − f ≤ ω = 1, from which we conclude that ω2(f) ≥ 0. Hence ω2 is a<br />

positive continuous linear functional on C(K), so by Lemma 1.79 ω2 = ω2(1) = 1,<br />

i.e. ω2 ∈ S(C(K)). It follows that<br />

F = {µ ∈ S(C(K)) : µ|Aff K = ω}<br />

is not empty. Note that F is a closed convex subseteq of S(C(K)). By the Krein-<br />

Milman theorem there is an extreme point ν in F. If ν = tµ1 + (1 − t)µ2, where<br />

t ∈]0, 1[ and µ1, µ2 ∈ S(C(K)), then ω = ν|Aff K = tµ1|Aff K + (1 − t)µ2|Aff K. Since<br />

µ1|Aff K, µ2|Aff K ∈ S(Aff K), the fact that ω is an extreme point in S(Aff K) implies<br />

that µ1|Aff K = µ2|Aff K = ω. Hence µ1, µ2 ∈ F, and since ν is an extreme point<br />

in F, we conclude that ν = µ1 = µ2. This shows that ν is in fact an extreme<br />

point in S(C(K)), not only in F. By Theorem 4.1 there is a point x ∈ K such<br />

that ν(f) = f(x), f ∈ C(K). In particular, ω(g) = ν(g) = g(x), g ∈ Aff K.<br />

Hence ω = ωx = Φ(x), i.e. ω ∈ Φ(K). Thus all extreme points in S(Aff K) are<br />

in Φ(K). Since Φ(K) is a compact convex subseteq of S(Aff K), it follows from<br />

the Krein-Milman theorem that Φ(K) = S(Aff K), i.e. Φ is surjective. As is wellknown<br />

a continuous bijection between compact Hausdorff spaces is automatically a<br />

homeomorphism, so the proof is complete. <br />

By Theorem 4.3 the real Banach space Aff K contains enough structure to recapture<br />

the compact convex set K defining it. In this way we can study K through


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 97<br />

the space Aff K. In order to use this viewpoint effectively, it is necessary to have an<br />

abstract characterization of the real Banach spaces which occur in this way.<br />

A convex cone in a real vector space E is a subset C ⊆ E such that 0 ∈ C and<br />

t1, t2 ∈ R + , x1, x2 ∈ C ⇒ t1x1 + t2x2 ∈ C. Alternatively, C is a convex cone if and<br />

only if C is a convex subset of E such that t ∈ R + , x ∈ C ⇒ tx ∈ C.<br />

Definition 4.4. An ordered Banach space is a real Banach space X containing<br />

a closed convex cone X+.<br />

Any ordered Banach space carries a natural pre-order; namely, x ≤ y ⇔ y − x ∈<br />

X+. Then X+ = {x ∈ X : x ≥ 0} and X+ is therefore referred to as the positive<br />

cone in X.<br />

Definition 4.5. An order unit space is an ordered Banach space X with a<br />

distinguished element u = 0 such that<br />

{x ∈ X : x ≤ 1} = {x ∈ X : −u ≤ x ≤ u}.<br />

Example 4.6. i) Let A be a unital C ∗ -algebra. Then Asa = {a ∈ A : a ∗ = a} is<br />

an ordered Banach space with positive cone consisting of the positive elements, i.e.<br />

Asa+ = A+ = {aa ∗ : a ∈ A}.<br />

If a ∈ Asa, then by spectral theory (Theorem 1.35), a ≤ 1 if and only if −1 ≤ a ≤<br />

1. Hence Asa is an order unit space.<br />

ii) Let X be an ordered Banach space with the property that there is a constant<br />

K > 0 such that<br />

a ≤ x ≤ b ⇒ x ≤ K max{a, b} .<br />

(An ordered Banach space with this property is called normal .) If u is an interior<br />

point in X+, there is a norm · u which is equivalent to the given one and turns X<br />

into an order unit space with X+ as positive cone and u as order unit. This norm is<br />

xu = inf{t > 0 : −tu ≤ x ≤ tu}.<br />

Let us prove this. First we show that {t > 0 : −tu ≤ x ≤ tu} = ∅ for all x ∈ X.<br />

Indeed, since u is an interior point of X+ there is an ǫ > 0 such that the closed ball<br />

u ≤<br />

of radius ǫ centered at u is contained in X+. For any x ∈ X, we have that − x<br />

ǫ<br />

x ≤ x<br />

x<br />

u because u − (u ± ǫ ǫ x ) ≤ ǫ when x = 0. Hence xu is welldefined and<br />

xu ≤ 1<br />

x. If −su ≤ x ≤ su and −tu ≤ y ≤ tu, then −(t+s)u ≤ x+y ≤ (s+t)u,<br />

ǫ<br />

proving that x+yu ≤ xu + yu. So · u is indeed a seminorm on X such that<br />

xu ≤ 1<br />

ǫ x. The normality condition implies that x ≤ Kuxu for all x ∈ X,<br />

so · u and · are indeed equivalent norms.<br />

It follows that most ”reasonable” ordered Banach spaces can be re-normed to<br />

become an order unit space.<br />

iii) If K is a compact convex set, Aff K is a real Banach space and<br />

is a closed convex cone. Since<br />

Aff K+ = {f ∈ Aff K : f(x) ≥ 0, x ∈ K}<br />

f = sup |f(x)| = inf{t > 0 : −t1 ≤ f ≤ t1},<br />

x∈K<br />

we see that Aff K is an order unit space with the constant function 1 as order unit.


98 1. FUNDAMENTALS<br />

We now show that Example 4.6 iii) gives the most general example of an order<br />

unit space, showing that this notion is the right abstraction of Aff K.<br />

Let X be an order unit space with order unit u. Set<br />

S(X) = {ω ∈ X ∗ : ω = ω(u) = 1}.<br />

Then S(X) is a convex set which is compact in the weak ∗ -topology by the Banach-<br />

Alaoglu theorem. Note that if X is the selfadjoint part of the unital C ∗ -algebra A, cf.<br />

Example 4.6 i), then S(X) ≃ S(A), i.e. there is a homeomorphism f : S(X) → S(A)<br />

such that f(tµ1 +(1 −t)µ2) = tf(µ1)+(1 −t)f(µ2), t ∈ [0, 1], µ1, µ2 ∈ S(X). (Such<br />

a map will be called an affine homeomorphism and we say that S(X) and S(A) are<br />

affinely homeomorphic.)<br />

Lemma 4.7. S(X) = {ω ∈ X ∗ : ω(x) ≥ 0, x ∈ X+, ω(u) = 1}.<br />

Proof. ⊆ : Let ω ∈ S(X). When x ≥ 0 and x = 0, 0 ≤ x ≤ u and hence<br />

x<br />

u − x<br />

x<br />

≤ 1 so that ω(u − ) ≤ 1. It follows that ω(x) ≥ 0. ⊇ : Assume that<br />

x x<br />

ω(x) ≥ 0, x ∈ X+, and ω(u) = 1. If x ≤ 1 we have that −u ≤ x ≤ u and hence<br />

−1 ≤ ω(x) ≤ 1, proving that ω ≤ 1, i.e. ω ∈ S(X). <br />

Lemma 4.8. For x ∈ X,<br />

x = sup{|ω(x)| : ω ∈ S(X)}.<br />

Proof. If x = 0 there is nothing to prove, so assume that x = 0. For any<br />

0 < s < x we have that either su − x or x + su is not in X+. If su − x /∈ X+,<br />

Hahn-Banach separation gives us an element ω ∈ X∗ such that ω(y) ≥ 0, y ∈ X+,<br />

and ω(su − x) < 0. Then µ = ω ∈ S(X) by Lemma 4.7. (ω(u) = 0 since ω is<br />

ω(u)<br />

non-zero and non-negative on X+.) Note that s < µ(x). If instead x + su /∈ X+<br />

we get in the same way an element µ ∈ S(X) such that s < −µ(x). It follows<br />

that s < sup{|µ(x)| : µ ∈ S(X)}. Since 0 < s < x was arbitrary, we see that<br />

x ≤ sup{|µ(x)| : µ ∈ S(X)}. The reversed inequality is trivial. <br />

Lemma 4.9. Let X be an order unit space. Then<br />

{ω ∈ X ∗ : ω ≤ 1} = co(S(X) ∪ −S(X)).<br />

Proof. Recall that co(S(X) ∪ −S(X)) = {tω + (1 − t)µ : ω, µ ∈ S(X) ∪<br />

−S(X), t ∈ [0, 1]} = {tω − (1 − t)µ : ω, µ ∈ S(X), t ∈ [0, 1]}. Define µ : [0, 1] ×<br />

S(X)×S(X) → X ∗ such that µ(t, µ, ν) = tµ−(1−t)ν. Then µ ([0, 1] × S(X) × S(X)) =<br />

co(S(X)∪−S(X)), so we conclude that co(S(X)∪−S(X)) is compact in the weak ∗ -<br />

topology because S(X) is and µ is continuous. Assume to get a contradiction that<br />

there is an element q ∈ {ω ∈ X ∗ : ω ≤ 1}\co(S(X) ∪ −S(X)). By the Hahn-<br />

Banach separation theorem there is then a weak ∗ -continuous functional which separates<br />

q and co(S(X) ∪ −S(X)). But a weak ∗ -continuous functional is given by<br />

an element of X, so we see that there is an element x ∈ X and a α ∈ R such<br />

that q(x) > α and µ(x) ≤ α, µ ∈ co(S(X) ∪ −S(X)). Note that α ≥ 0 and that<br />

|µ(x)| ≤ α for all µ ∈ S(X). By Lemma 4.8 it follows that x ≤ α, which is<br />

impossible since q(x) > α. This proves the lemma. <br />

For x ∈ X, define fx ∈ Aff S(X) by<br />

fx(ω) = ω(x), ω ∈ S(X).


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 99<br />

Theorem 4.10. The map x ↦→ fx is an isometric isomorphism of X onto<br />

Aff S(X) which takes X+ onto Aff S(X)+ and u to the constant function 1.<br />

Proof. Set Φ(x) = fx. Then Φ is a linear map which is an isometry by Lemma<br />

4.8. It follows from Lemma 4.7 that Φ(X+) ⊆ Aff S(X)+. It remains only to show<br />

that Φ(X+) = Aff S(X)+. Since Φ is an isometry and X+ is closed in X, Φ(X+) is<br />

a closed convex cone in Aff S(X) and if Φ(X+) is not all of Aff S(X)+ the Hahn-<br />

Banach separation theorem gives us elements g ∈ Aff S(X)+, ν ∈ (Aff S(X)) ∗ and<br />

β ∈ R such that ν(z) ≤ β for all z ∈ Φ(X+) and ν(g) > β. Note first that β ≥ 0<br />

since 0 ∈ Φ(X+) and that ν(z) ≤ 0 for all z ∈ Φ(X+) since Φ(X+) is a cone. Thus<br />

while<br />

ν(g) > 0 (4.1)<br />

ν ◦ Φ(z) ≤ 0 (4.2)<br />

for all z ∈ X+. It follows from Lemma 4.9 that ν = t+ν+ − t−ν− where t± ≥ 0 and<br />

ν± ∈ S (Aff S(X)). Note that ν± are given by evaluation at two states ω± ∈ S(X)<br />

by Theorem 4.3. Then (4.2) implies that t+ω+(z) = t+ν+ ◦ Φ(z) ≤ t−ν− ◦ Φ(z) =<br />

t−ω−(z) for all z ∈ X+. In combination with (4.1) this implies in particular that<br />

t− = 0. Thus there is a γ ∈ S(X) and some t ≥ 0 such that<br />

ω− = t+<br />

t−<br />

ω+ + tγ.<br />

Since g is affine this implies that g (ω−) = t+<br />

t− g (ω+) + tg(γ), and hence that<br />

t−ν−(g) = t+ν+(g) + t−tg(γ).<br />

Thus ν(g) = t+ν+(g) − t−ν−(g) = −t−tg(γ). Note that −t−tg(γ) ≤ 0 since<br />

g ∈ Aff S(X)+ so that we have contradicted (4.1). Hence Φ(X+) must be all of<br />

Aff S(X)+. <br />

Corollary 4.11. Let X be an order unit space with order unit u. For x ∈ X,<br />

x ∈ X+ ⇔ xu − x ≤ 1 .<br />

We can now pass from compact convex sets to order unit spaces and from order<br />

unit spaces to compact convex sets. The functors are K → Aff K and X → S(X),<br />

respectively. By Theorem 4.3 and Theorem 4.10 the functors are inverses of each<br />

other, i.e. S(Aff K) = K by Theorem 4.3 and AffS(X) = X by Theorem 4.10. We<br />

shall later see that this equivalence of functors may be considered as an extension of<br />

the equivalence between compact Hausdorff spaces and unital abelian C ∗ -<strong>algebras</strong>.<br />

Let X be a compact metric space. A Borel measure µ on X is a probability<br />

measure when µ(X) = 1. The set of Borel probability measures on X will be<br />

denoted by M1(X). Since we assume that X is metrizable a Borel probability<br />

measure µ ∈ M1(X) is regular, i.e. satisfies that<br />

µ(B) = sup{µ(C) : C ⊆ B, C closed} = inf{µ(U) : B ⊆ U, U open}<br />

for all Borel subsets B of X. If X is not metrizable these properties will have to<br />

be added to the definition. There is a one-to-one correspondance between the Borel<br />

probability measures on X and the states of C(X) given by integration with respect<br />

to measures (the direction ’from measures to states’) and the Riesz representation


100 1. FUNDAMENTALS<br />

theorem (the direction ’from states to measures’). Thus the state sµ on C(X) given<br />

by the measure µ ∈ M1(X) is<br />

<br />

sµ(f) = f dµ, f ∈ C(X),<br />

and the Borel probability measure µs corresponding to the state s ∈ S(C(X)) is<br />

given by<br />

µs(V ) = sup{s(f) : f ∈ C(X), 0 ≤ f ≤ 1, supp(f) ⊆ V }<br />

when V ⊆ X is open.<br />

By using the identification M1(X) = S(C(X)), the space M1(X) of Borel probability<br />

measures becomes a compact convex set in the weak ∗ -topology. Note that the<br />

convex structure is the natural one where (tµ1+(1−t)µ2)(B) = tµ1(B)+(1−t)µ2(B)<br />

for all Borel sets B. By Theorem 4.1 and Exercise 4.2 the extreme points of the<br />

compact convex set M1(X) are exactly the point measures, i.e. measures δx obtained<br />

by fixing an x ∈ X and setting<br />

<br />

0, x /∈ B<br />

δx(B) =<br />

1, x ∈ B,<br />

for every Borel set B ⊆ X. Note that f dδx = f(x), f ∈ C(X).<br />

Exercise 4.12. 1) Show that a compact Hausdorff space X is metrizable<br />

if and only if C(X) is separable.<br />

(Sketch of a possible solution : If d is a metric, show that there is a<br />

sequence {Bn}n∈N of balls in X such that dist(·, Bn), n ∈ N, separates the<br />

points of X. By the Stone-Weierstrass theorem the complex polynomials in<br />

these functions, all of whose coefficients have rational real and imaginary<br />

parts, will be dense in C(X). - Conversely, if {fn} is a dense sequence in<br />

the unit ball of C(X), consider d(x, y) = ∞<br />

n=1 2−n |fn(x) − fn(y)|.)<br />

2) Show that a subseteq of a separable normed vector space is itself separable.<br />

3) Let K be a compact convex set. Show that K is metrizable if and only if<br />

Aff K is separable.<br />

4) Let X be a metrizable compact Hausdorff space. Show that M1(X) is<br />

metrizable.<br />

Consider now a compact convex set K.<br />

Lemma 4.13. 1) For every Borel probability measure µ ∈ M1(K) there is<br />

a unique point b(µ) ∈ K such that<br />

<br />

f(b(µ)) = f dµ, f ∈ Aff K.<br />

2) b : M1(K) → K is an affine continuous surjection.<br />

Proof. 1) follows immediately from Theorem 4.3. 2) : When µ1, µ2 ∈ M1(K), t ∈<br />

[0, 1], then f d(tµ1+(1−t)µ2) = t f dµ1+(1−t) f dµ2 for all f ∈ C(K). In particular,<br />

it holds for all f ∈ Aff K and hence b(tµ1 +(1−t)µ2) = tb(µ1)+(1−t)b(µ2),<br />

i.e. b is affine. When {µα} is a net in M1(K) converging to µ ∈ M1(K), we<br />

have in particular that limα f dµα = f dµ for all f ∈ Aff K, and hence that<br />

limα f(b(µα)) = f(b(µ)) for all f ∈ Aff K. By Theorem 4.3 this implies that<br />

limα b(µα) = b(µ).


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 101<br />

The point b(µ) ∈ K will be called the barycenter of µ.<br />

Since the point measures {δx : x ∈ K} are the extreme points of M1(K), we know<br />

from the Krein-Milman theorem that every meausure µ ∈ M1(K) is the limit of a<br />

sequence (or net if K is not metrizable) of convex combinations of point measures.<br />

We shall need the fact that the elements in this approximating sequence (or net)<br />

can be chosen such that all of them have the same barycenter as µ.<br />

We say that µ ∈ M1(K) is supported on the Borel set B ⊆ K when µ(B) = 1.<br />

More generally, we say that a positive Borel measure ν is supported on B when<br />

ν(K\B) = 0. Note that a measure in M1(K) is a (finite) convex combination of<br />

point measures if and only if it is supported on a finite subseteq of K. Hence the<br />

result which we shall need can be stated as follows.<br />

Lemma 4.14. Let µ ∈ M1(K). There is then a net {να} in M1(K) such that<br />

each να is supported on a finite set, b(να) = b(µ) for all α and limα να = µ.<br />

Proof. We will define a directed set as follows. The elements of the set consists<br />

of finite partitions U = {Ui : i = 1, 2, . . ., n} of K into Borel sets, i.e. each Ui is a<br />

Borel set, Ui∩Uj = ∅ when i = j and <br />

i Ui = K. We order the set of Borel partitions<br />

in the following way : When U = {Ui} and V = {Vi} are two such partitions we<br />

write U ≤ V when every Vi is a subseteq of some Uj, i.e. when V is a refinement of<br />

U. In this way the Borel partitions form a directed set. Next we associate a finitely<br />

supported measure µU to each Borel partition in the following way. First choose<br />

a fixed, but arbitrary meausure µ0 ∈ M1(K). When U = {Ui : i = 1, 2, . . ., n} is<br />

a Borel partition, we set µi = µ0 when µ(Ui) = 0, and when µ(Ui) = 0 we define<br />

µi ∈ M1(K) as the unique Borel probability measure such that<br />

<br />

f dµi = 1<br />

µ(Ui)<br />

<br />

f1Ui<br />

dµ, f ∈ C(K).<br />

The existence and uniqueness of such a µi follows from the Riesz representation<br />

theorem. Set xi = b(µi) for all i and define<br />

n<br />

µU = µ(Ui)δxi .<br />

i=1<br />

Then µU has finite support (it is supported on {x1, x2, . . ., xn}) and we have that<br />

<br />

n<br />

n<br />

<br />

n<br />

<br />

f dµU = µ(Ui)f(xi) = µ(Ui) f dµi = f1Ui dµ = f dµ<br />

i=1<br />

i=1<br />

for all f ∈ Aff K, from which we conclude that b(µU) = b(µ). We assert that the net<br />

{µU} converges to µ. To prove this we must show that limU f dµU = f dµ for all<br />

f ∈ C(K). So fix such an f ∈ C(K) and use the compactness of K to find an open<br />

convex cover Wi, i = 1, 2, . . ., m, of K such that x, y ∈ Wi ⇒ |f(x)−f(y)| < ǫ for all<br />

i. Set U1 = W1 and Ui = Wi\ <br />

j


102 1. FUNDAMENTALS<br />

when λj = 0. To prove this we first prove that<br />

xj is in the closed convex hull of Vj. (4.4)<br />

If namely, (4.4) was not true the Hahn-Banach separation theorem would give us a<br />

g ∈ Aff K and a α ∈ R such that g(xj) > α while g(x) ≤ α for all x ∈ Vj. This would<br />

imply that λ −1 <br />

g1Vj j dµ ≤ αλ−1 1Vj j dµ = α in contradiction with the estimate<br />

λ −1 <br />

g1Vj j dµ = g dµj = g(b(µj)) = g(xj) > α. Next (4.4) is used to establish<br />

(4.3) as follows. Since U ≤ V there is a k such that Vj ⊆ Uk ⊆ Wk. Since we chose<br />

the Wi’s to be convex, we conclude from (4.4) that xj ∈ Wk. When x ∈ Vj ⊆ Wk<br />

we see from the choice of Wk that |f(x) − f(xj)| ≤ ǫ, proving (4.3). By using (4.3)<br />

we can now estimate<br />

<br />

n<br />

n<br />

<br />

| f dµV − f dµ| = | λif(xi) − f1Vi dµ|<br />

≤<br />

n<br />

<br />

λi|f(xi) −<br />

i=1<br />

i=1<br />

f dµi| =<br />

n<br />

i=1<br />

<br />

λi|<br />

i=1<br />

(f(xi) − f)1Vi dµi| ≤<br />

n<br />

λiǫ = ǫ.<br />

Thus we have shown that limU µU = µ as asserted. <br />

The reader can check on any compact convex subseteq of R 2 he or she can draw,<br />

that any point of the set can be realized as the barycenter of a Borel probability<br />

measure which is supported on the extreme points. This holds in general for all<br />

compact convex sets, except for the annoying fact that the set ∂eK of extreme<br />

points of K need not be a Borel set when K is not metrizable. Our next goal is to<br />

prove this in the metrizable case.<br />

Set<br />

S(K) = {f ∈ CR(K) : f(tx + (1 − t)y) ≤ tf(x) + (1 − t)f(y), t ∈ [0, 1], x, y ∈ K},<br />

i.e. S(K) consists of the continuous convex functions on K. For any f ∈ CR(K) we<br />

define the upper envelope f by<br />

f(x) = inf{g(x) : −g ∈ S(K), g ≥ f}.<br />

A function g ∈ CR(K) is concave when −g is convex, so the upper envelope of f<br />

is the infimum over all continuous concave functions dominating f. In particular, it<br />

follows straightforwardly that f is concave and upper semicontinuous. (Recall that<br />

a function h : K → R is upper semicontinuous when h −1 (] − ∞, α[) is open for all<br />

α ∈ R.) Therefore f is Borel measurable.<br />

Lemma 4.15. Let f ∈ CR(K). Then<br />

is the closed convex hull of<br />

T = {(ω, t) ∈ K × R : t ≤ f(ω)}<br />

S = {(ω, t) ∈ K × R : t ≤ f(ω)}.<br />

Proof. Note that T is convex since f is concave and that T is closed since f<br />

is upper semicontinuous. Since f ≤ f, it is clear that S ⊆ T so we have almost for<br />

free that co(S) ⊆ T. Assume that there is a point (ω0, t0) ∈ T \co(S). If so we may<br />

i=1


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 103<br />

apply the Hahn-Banach separation theorem to get a continuous linear functional g<br />

on (Aff K × R) ∗ and a real number α such that<br />

g(ω0, t0) > α and g(ω, t) ≤ α when t ≤ f(ω)<br />

Now g is of the form g(ω, t) = h(ω) + tλ for some weak ∗ -continuous functional h<br />

on Aff K ∗ and some λ ∈ R. Thus h(ω0) + t0λ > α and h(ω) + f(ω)λ ≤ α for all<br />

ω ∈ K. Inserting ω = ω0 and subtracting we get that λ(f(ω0) − t0) < 0 which<br />

implies that λ > 0 because t0 > f(ω0). Since λ > 0 we can now conclude that<br />

k(ω) = λ −1 (α − h(ω)) defines an element of Aff K such that k ≥ f. In particular,<br />

k ≥ f and t0 > k(ω0) ≥ f(ω0), contradicting that (ω0, t0) was an element of T. <br />

Lemma 4.16. Assume that K is metrizable. For every f ∈ CR(K),<br />

∂eK ⊆ {ω ∈ K : f(ω) = f(ω)}.<br />

Proof. Assume that f(ω) > f(ω). Since (ω, f(ω)) is in the set T of Lemma<br />

4.15, there is a sequence {ηn} of finite convex combinations, ηn = Nn<br />

i=1 tn i (ω n i , s n i ),<br />

where s n i ≤ f(ωn i ) for all i, n, such that limn→∞ ηn = (ω, f(ω)). By passing to<br />

a subsequence we may assume that ν = limn→∞<br />

) = limn<br />

follows that b(ν) = limn b( Nn<br />

i=1 tn i δω n i<br />

<br />

f dν = lim n<br />

<br />

Nn <br />

f d(<br />

i=1<br />

t n i δω n i ) = lim n<br />

Nn i=1 tni δωn i exists in M1(K). It<br />

Nn i=1 tni ωn i = ω. Furthermore,<br />

Nn <br />

t<br />

i=1<br />

n i f(ωn i ) ≥ lim n<br />

Nn <br />

t n i sni i=1<br />

= f(ω),<br />

from which we see that f dν > f(ω). Now, by Lemma 4.14, there is a sequence<br />

of finitely supported measures ρn ∈ M1(K) such that b(ρn) = b(ν) = ω for all n and<br />

limn ρn = ν. For some large enough N ∈ N we have that f dρN > f(ω). Since ρN<br />

is finitely supported there are points xi ∈ K and numbers ri ∈]0, 1], i = 1, 2, . . ., mN<br />

such that ρN = mN riδxi i=1 . Then<br />

mN <br />

ω = b(ρN) = rixi. (4.5)<br />

Since f(ω) < f dρN = mN<br />

i=1 rif(xi), we conclude that not all xi’s equal ω so that<br />

(4.5) is a non-trivial convex combination, and hence we see that ω /∈ ∂eK. This<br />

proves that ∂eK ⊆ {ω : f(ω) = f(ω)}. <br />

Proposition 4.17. Let K be a metrizable compact convex set. There is a convex<br />

function F ∈ S(K) such that<br />

i=1<br />

∂eK = {x ∈ K : F(x) = F(x)}.<br />

Proof. Since K is metrizable we can find a metric d on K defining the topology.<br />

In fact, since Aff K is separable by Exercise 4.12, we can define a metric for the<br />

topology by<br />

∞<br />

d(x, y) = 2 −n |fn(x) − fn(y)|<br />

n=1<br />

where {fn} is a dense sequence in the unit ball of Aff K. Note that with this<br />

definition the function d(x, · ) is convex for all x ∈ K. Since K is metrizable and


104 1. FUNDAMENTALS<br />

compact it is also separable, so we can find a dense sequence {xn} in K. Since<br />

sup x,y∈K d(x, y) < ∞, the sum<br />

F(x) =<br />

∞<br />

n=1<br />

2 −n d(xn, x) 2<br />

converges uniformly in x ∈ K. Since t ↦→ t 2 is convex on R + it follows that F ∈<br />

S(K). Let ω /∈ ∂eK. We assert that F(ω) > F(ω). To see this note that ω =<br />

1<br />

2 (µ1 + µ2) for some µ1 = µ2 in K since ω /∈ ∂eK. There is an n ∈ N such that<br />

d(xn, µ1) < d(xn, µ2). Then<br />

d(xn, ω) 2 ≤ 1<br />

2 (d(xn, µ1) + d(xn, µ2) 2<br />

= 1<br />

2 (d(xn, µ1) 2 + d(xn, µ2) 2 ) − 1<br />

4<br />

< 1<br />

2 (d(xn, µ1) 2 + d(xn, µ2) 2 ),<br />

d(xn, µ1) − d(xn, µ2) 2<br />

from which it follows that F(ω) < 1<br />

2 (F(µ1)+F(µ2)). Thus, if g ∈ S(K) and −g ≥ F,<br />

then<br />

−g(ω) = −g( 1<br />

2 (µ1 + µ2)) ≥ 1<br />

2 (−g(µ1) − g(µ2)) ≥ 1<br />

2 (F(µ1) + F(µ2)) = F(ω) + δ,<br />

where δ = 1<br />

2 (F(µ1) + F(µ2)) − F(ω). By taking the infimum over all such g we<br />

see that F(ω) ≥ F(ω) + δ, so that F(ω) > F(ω). We can now conclude that<br />

{ω ∈ K : F(ω) = F(ω)} ⊆ ∂eK. The reversed inclusion follows from Lemma 4.16.<br />

<br />

It follows from Proposition 4.17 that the extreme boundary ∂eK of a metrizable<br />

convex set K is the intersection of a countable number of open subseteqs, and hence,<br />

in particular, is a Borel set. Indeed, (F − F) −1 (] − ∞, α[) = <br />

β∈R F −1 (] − ∞, α +<br />

β[) ∩ F −1 (]β, ∞[) is open for all α ∈ R and<br />

∂eK = <br />

(F − F) −1 (] − ∞, 1<br />

n [)<br />

by Proposition 4.17.<br />

n∈N<br />

Lemma 4.18. S(K) − S(K) is dense in CR(K), i.e. every continuous realvalued<br />

function on K can be approximated uniformly by the difference between two<br />

continuous convex functions.<br />

Proof. By Theorem 4.3 Aff K separates the points of K. By the real Stone-<br />

Weierstrass theorem it follows that polynomials in elements from Aff K form a dense<br />

subalgebra of CR(K), i.e. every f ∈ CR(K) can be approximated arbitrarily closely<br />

(in the uniform norm) by a function of the form K ∋ ω → P(g1(ω), g2(ω), . . ., gn(ω))<br />

for some polynomial P of n real variables. Set L = g1(K)×g2(K)×· · ·×gn(K) ⊆ R n .<br />

It is wellknown that a smooth function h : [0, 1] → R is convex if and only if<br />

h ′′ (t) ≥ 0 for all t ∈ [0, 1]. A polynomial Q : L → R is convex on L if and only<br />

if t ↦→ Q(tx + (1 − t)y) is convex on [0, 1] for all x, y ∈ L, so we conclude that<br />

Q is convex if and only if <br />

i,j<br />

∂ 2 Q<br />

∂xi∂xj (tx + (1 − t)y)(xi − yi)(xj − yj) ≥ 0 for all


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 105<br />

x, y ∈ L, t ∈ [0, 1]. In particular, it suffices that the selfadjoint n × n-matrix<br />

∂2Q (χ)<br />

∂xi∂xj<br />

<br />

is positive in Mn(C) for all χ ∈ L. Since χ ↦→ ∂2Q ∂xi∂xj (χ) is continuous, there is a<br />

constant K ∈ R + such that<br />

2K + ∂2Q (χ)<br />

∂xi∂xj<br />

<br />

is positive in Mn(C) for all χ ∈ L. Set P1(x1, x2, . . .,xn) = P(x1, x2, . . .,xn) +<br />

K n<br />

i=1 x2 i and P2(x1, x2, . . .,xn) = K n<br />

i=1 x2 i . Then P1 and P2 are both convex<br />

polynomials on L and ω ↦→ P1(g1(ω), . . ., gn(ω)) and ω ↦→ P2(g1(ω), . . ., gn(ω)) are<br />

convex functions on K whose difference is ω ↦→ P(g1(ω), . . ., gn(ω)). <br />

Theorem 4.19. Let K be a metrizable compact convex set. Then every point<br />

x ∈ K is the barycenter of a Borel probability measure supported on the extreme<br />

boundary, ∂eK, of K.<br />

Proof. We introduce a partial ordering < on M1(K) by µ < ν iff f dµ ≤<br />

f dν, f ∈ S(K). This relation is obviously reflexive (µ < µ) and transitive<br />

(µ < ν, ν < δ ⇒ µ < δ). That it is also anti-symmetric follows from Lemma<br />

4.18 : If µ < ν and ν < µ, then clearly g dµ = g dν for all g ∈ S(K), and<br />

so f dµ = f dν for all f ∈ CR(K) by Lemma 4.18, i.e. µ = ν. Consider<br />

L = {µ ∈ M1(K) : b(µ) = x} with the partial ordering < . Note that δx ∈ L<br />

so that L is a non-empty partially ordered set. If G is a totally ordered subseteq<br />

of L we have immediately that limµ∈G g dµ exists in R for all g ∈ S(K). It<br />

follows from Lemma 4.18 that limµ∈G f dµ exists for all f ∈ CR(K), so there is a<br />

ν ∈ M1(K) such that limG µ = ν. By the continuity of the barycenter map, we have<br />

that b(ν) = limG b(µ) = x, i.e. ν ∈ L. Clearly, µ < ν for all µ ∈ G, so that ν is<br />

an upper bound for G. We can now apply Zorns lemma to conclude that there is a<br />

maximal element µ in L. We use µ to define a Minkowski functional p on CR(K) by<br />

<br />

p(f) = f dµ, f ∈ CR(K).<br />

Since tf = tf, f + g ≤ f + g for all f, g ∈ CR(K), t ≥ 0, the linearity of<br />

the integral yields that p is a Minkowski functional. Let F ∈ S(K) such that<br />

∂eK = {ω ∈ K : F(ω) = F(ω)}, cf. Proposition 4.17. By the Hahn-Banach theorem<br />

there is a linear functional l : CR(K) → R such that l(F) = p(F) and l(f) ≤ p(f) for<br />

all f ∈ CR(K). If g ∈ CR(K) and g ≤ 0, then g ≤ 0 and l(g) ≤ p(g) = g dµ ≤ 0,<br />

proving that l is a positive linear functional. By the Riesz representation theorem<br />

there is a positive Borel measure ν on K such that l(f) = f dν, f ∈ CR(K).<br />

When g ∈ S(K), −g is concave and hence<br />

<br />

<br />

−l(g) = −g dν ≤ p(−g) =<br />

<br />

−g dµ = −<br />

g dµ. (4.6)<br />

Since any affine function h ∈ Aff K is both convex and concave, (4.6) implies that<br />

h dν = h dµ for all h ∈ Aff K. In particular, ν(K) = 1 dν = 1 dµ = 1,<br />

showing that ν is a Borel probability measure. Also it follows that b(ν) = b(µ) = x,<br />

i.e. ν ∈ L. Now (4.6) shows that µ < ν. By the maximality of µ we must have<br />

that µ = ν and hence F dµ = p(F) = l(F) = F dν = F dµ. It follows


106 1. FUNDAMENTALS<br />

from this that µ must be supported on ∂eK, i.e. that µ(∂eK) = 1. Indeed, if<br />

Ωn = {ω ∈ K : F(ω) − F(ω) ≥ 1 } had positive measure for some n, we could<br />

n<br />

conclude that F − F dµ ≥ <br />

1 F − F dµ ≥ Ωn n µ(Ωn) > 0, which is impossible.<br />

Therefore K\∂eK = <br />

n Ωn has measure 0, i.e. µ(∂eK) = 1. <br />

4.2. Choquet simplices. Among the compact convex sets there is a special<br />

class with particularly nice properties - the Choquet simplices, or just simplices.<br />

Definition 4.20. A compact convex set K is a Choquet simplex when Aff K<br />

has the Riesz decomposition property in the order given by the cone Aff K+ = {f ∈<br />

Aff K : f(x) ≥ 0, x ∈ K}.<br />

Note that by Proposition 3.39 we could equally well require Aff K to have the<br />

Riesz interpolation property. In Section 4.1 we saw that the compact convex sets<br />

are in one to one correspondance with the order unit spaces. By definition the state<br />

space of an order unit space is a Choquet simplex if and only if the order unit space<br />

has the Riesz interpolation (or decomposition) property.<br />

There are other ways to define a Choquet simplex (e.g. as the base of a lattice<br />

cone), but we prefer the given one because it emphasizes the property which makes<br />

the simplices so important in connection with dimension groups. As it stands,<br />

however, it is not so easy to identify a Choquet simplex among other compact<br />

convex sets, so we seek a more geometric condition for a compact convex set to be<br />

a Choquet simplex. We shall only be interested in compact convex sets which are<br />

metrisable, and since the theory presents additional difficulties in the non-metrisable<br />

case we shall simply confine ourself to the metrizable case whenever it is convenient.<br />

Example 4.21. Let X be a compact metric space. M1(X) is a metrizable Choquet<br />

simplex. To see this, define a map Φ : CR(X) → Aff M1(X) by<br />

<br />

Φ(f)(µ) = f dµ, µ ∈ M1(X).<br />

Φ is then clearly a linear map which is easily seen to be an isometry such that<br />

f ≥ 0 ⇔ Φ(f) ≥ 0. If h ∈ (Aff M1(X))+, the function<br />

X ∋ x ↦→ h(δx)<br />

is continuous so it defines an element f of CR(X) which is clearly non-negative. Since<br />

Φ(f)(δx) = f(x) = h(δx), x ∈ X, we see that Φ(f) and h agree on the extremepoints<br />

of M1(X). Since h and Φ(f) are both continuous and affine, it follows from the<br />

Krein-Milman theorem that h = Φ(f). Hence Aff M1(X) is isomorphic to CR(X)<br />

as partially ordered vector spaces. Since CR(X) is obviously a lattice , i.e. the<br />

supremum of two elements exists in CR(X), cf. Definition 4.22 below, it follows that<br />

Aff M1(X) is also a lattice. In particularly, Aff M1(X) has the Riesz interpolation<br />

property, so M1(X) is a Choquet simplex. The fact that M1(X) is metrizable is left<br />

as an exercise, cf. Exercise 4.12.<br />

This example comprises a large stock of examples of Choquet simplices, including<br />

the classical simplices<br />

∆n = {(t1, t2, . . .,tn) ∈ [0, 1] n n<br />

: ti = 1},<br />

n ∈ N. Indeed, if F is a finite set consisting of n elements, then M1(F) is affinely<br />

homeomorphic to ∆n as the reader can easily verify.<br />

i=1


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 107<br />

The next goal here will be to prove that ’the way a point is expressed as a convex<br />

combination of extreme points is unique if and only if the set is a simplex’.<br />

Definition 4.22. An ordered Banach space X is a lattice when any pair x1, x2 ∈<br />

X have a supremum in X; i.e. if there is an element x1 ∨ x2 ∈ X such that<br />

xi ≤ x1 ∨ x2, i = 1, 2, and such that xi ≤ y, i = 1, 2 ⇒ x1 ∨ x2 ≤ y.<br />

Hence X is a lattice if and only if any finite set has a supremum as well as an<br />

infimum. It is wellknown that CR(Y ) is a lattice (in the natural ordering) for any<br />

compact Hausdorff space Y . What may be less obvious is that the dual space of<br />

CR(Y ) is also a lattice. To make this precise we return for a moment to general<br />

setup where X is an order unit space with order unit u and positive cone X+. We<br />

turn the dual space X ∗ into an ordered Banach space by setting<br />

X ∗ + = {l ∈ X ∗ : l(X+) ⊆ [0, ∞[}.<br />

It follows from Lemma 4.9 that X∗ + − X∗ + = X∗ , so X∗ becomes an ordered Banach<br />

space in this way. In fact, since X+ − X+ = X, X∗ + is a proper cone in the sense<br />

that X ∗ + ∩ (−X∗ + ) = {0}. When the dual space X∗ is considered as an ordered<br />

Banach space in this way, we refer to X ∗ as the ordered dual of X.<br />

Lemma 4.23. If X is an order unit space with the Riesz interpolation property,<br />

then the ordered dual X ∗ is a lattice.<br />

Proof. Let l1, l2 ∈ X ∗ . For x ∈ X+ we define l1 ∨ l2(x) ∈ R by<br />

l1 ∨ l2(x) =<br />

n<br />

sup{ max{l1(ai), l2(ai)} : x = a1 + a2 + · · · + an, ai ∈ X+, i = 1, 2, . . ., n}.<br />

i=1<br />

Note that by Lemma 4.9 we have that li = l + i − l− i where l+ i , l− i ∈ X∗ + and l + i<br />

li, l −<br />

i ≤ li, i = 1, 2. Thus<br />

n<br />

max{l1(ai), l2(ai)} ≤<br />

n<br />

l + 1 (ai) + l − 1 (ai) + l + 2 (ai) + l − 2 (ai)<br />

i=1<br />

i=1<br />

= l + 1 (x) + l − 1 (x) + l + 2 (x) + l − 2 (x) ≤ 2x(l1 + l2))<br />

whenever ai ∈ X+ and <br />

i ai = x. Therefore the supremum occuring in (4.23) exists.<br />

It is (almost) obvious that l1 ∨ l2 is superadditive on X+, i.e. that l1 ∨ l2(x + y) ≥<br />

l1 ∨ l2(x) + l1 ∨ l2(y), x, y ∈ X+. To prove the reversed inequality, we shall use<br />

the Riesz decomposition property. Indeed, if x + y = a1 + a2 + · · · + an, where<br />

ai ∈ X+ for all i, then we may write ai = bi + ci where 0 ≤ bi, 0 ≤ ci for all i and<br />

<br />

i bi = x, <br />

i ci = y, we find that<br />

n<br />

max{l1(ai), l2(ai)} =<br />

i=1<br />

≤<br />

n<br />

max{l1(bi) + l1(ci)), l2(bi) + l2(ci)}<br />

i=1<br />

n<br />

max{l1(bi), l2(bi)} + max{l1(ci), l2(ci)} ≤ l1 ∨ l2(x) + l1 ∨ l2(y).<br />

i=1<br />

Since the decomposition x + y = <br />

i ai into positive elements was arbitrary, we<br />

conclude that l1 ∨ l2(x + y) ≤ l1 ∨ l2(x) + l1 ∨ l2(y). Hence l1 ∨ l2 is additive<br />

on X+ and we can extend it to an additive function l1 ∨ l2 : X → R by setting<br />


108 1. FUNDAMENTALS<br />

l1 ∨l2(x−y) = l1 ∨l2(x)−l1 ∨l2(y) when x, y ∈ X+. As observed above, we have that<br />

|l1 ∨l2(x)| ≤ 2x(l1+l2)) when x ∈ X+. In general, x = a−b, where a, b ∈ X+<br />

and a ≤ x and b ≤ x, so in general we have that |l1 ∨ l2(x)| ≤ 4x(l1 +<br />

l2)), x ∈ X. Hence l1 ∨ l2 is continuous. Furthermore, l1 ∨ l2(nx) = nl1 ∨ l2(x)<br />

when n ∈ N, so it follows that l1 ∨ l2(qx) = ql1 ∨ l2(x) for all rational numbers q.<br />

The density of Q in R, combined with the continuity of l1 ∨ l2, implies that l1 ∨ l2<br />

is linear, i.e. l1 ∨ l2 ∈ X∗ . Clearly, li ≤ l1 ∨ l2, i = 1, 2, and if l ∈ X∗ satisfies<br />

that li ≤ l, i = 1, 2, then n i=1 max{l1(ai), l2(ai)} ≤ <br />

i l(ai) = l(x) whenever<br />

x, a1, a2, · · · , an ∈ X+ are such that x = a1 + a2 + · · · + an. Hence l1 ∨ l2(x) ≤ l(x),<br />

proving that l1 ∨ l2 is indeed the supremum of l1 and l2 in X∗ . <br />

Lemma 4.24. For every f ∈ CR(K) and ω ∈ K,<br />

n<br />

f(ω) = sup{ tif(ωi) : ti ∈ [0, 1], ωi ∈ K, <br />

ti = 1, <br />

tiωi = ω , n ∈ N}.<br />

i=1<br />

Proof. Since f is concave we have that<br />

n<br />

f(ω) ≥ tif(ωi) ≥<br />

i=1<br />

i<br />

n<br />

tif(ωi)<br />

whenever ti ∈ [0, 1], ωi ∈ K, <br />

inequalities making up the identity in the lemma. To prove the other, we use Lemma<br />

i=1<br />

i ti = 1, and <br />

i tiωi = ω. This proves one of the<br />

4.15 to find λα i ∈ [0, 1], ωα i ∈ K, tα i ∈ R and nα ∈ N such that nα i=1 λαi = 1, tα i ≤<br />

f(ωα i ) and<br />

nα <br />

lim λ<br />

α<br />

i=1<br />

α i ωα i = ω, lim nα<br />

λ<br />

α<br />

i=1<br />

α i tαi = f(ω).<br />

Set<br />

nα <br />

µα = λ<br />

i=1<br />

α i δωα i<br />

and pass to a weak ∗ convergent subseteq µα<br />

one has b(µ) = limα ′ b(µα ′) = ω and<br />

<br />

f dµ = lim<br />

α ′<br />

f dµα ′ ≥ lim α ′<br />

i<br />

′. If µ denotes the limit of this subnet<br />

n α ′<br />

<br />

i=1<br />

λ α′<br />

i tα′ i = f(ω).<br />

Let ǫ > 0. By Lemma 4.14 we can find a finitely supported measure ν ∈ M1(K) such<br />

that f dν > f dµ − ǫ and b(ν) = b(µ) = ω. Since ν is finitely supported there<br />

are ti ∈ [0, 1], ωi ∈ K, such that <br />

i ti = 1, and <br />

i tiδωi = ν. Since n i=1 tiωi =<br />

b(ν) = ω, and<br />

n<br />

<br />

tif(ωi) = f dν > f dµ − ǫ ≥ f(ω) − ǫ,<br />

i=1<br />

we get the second inequality. <br />

Lemma 4.25. Let g ∈ Aff K such that g(x) > 0 for all x ∈ K, and let p ∈ Aff K∗ .<br />

Then<br />

inf{q(g) : q ∈ (Aff K) ∗ + , q ≥ p} = sup{p(h) : 0 ≤ h ≤ g}.


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 109<br />

Proof. When q ≥ p and 0 ≤ h ≤ g, then p(h) ≤ q(h) ≤ q(g), proving that<br />

sup{p(h) : 0 ≤ h ≤ g} ≤ inf{q(g) : q ∈ (Aff K) ∗ + , q ≥ p}. Note that β = sup{p(h) :<br />

0 ≤ h ≤ g} is finite since p can be written as the difference between two positive<br />

functionals. Let ǫ > 0 and choose m ∈ N such that mǫ ≥ β − p(g). Define linear<br />

subspaces of Aff K ⊕ Aff K by<br />

C = {(x, −x) : x ∈ Aff K}, D = R(mg, g) + C .<br />

Note that D is a closed subspace of Aff K⊕Aff K and hence a real Banach space. Set<br />

D+ = D ∩ (Aff K+ × Aff K+) and note that (mg, g) is in the interior of D+ relative<br />

to D. Since (D, D+) is a normal ordered Banach space, we know from Example 4.6<br />

2) that there is a norm · ′ , equivalent to the one inherited from Aff K ⊕ Aff K,<br />

such that D is an order unit space with order unit (mg, g). Since g = 0 we have<br />

that (mg, g) /∈ C. We can therefore define a linear functional f on D so that<br />

f(t(mg, g) + (x, −x)) = tm(β + ǫ) + p(x)<br />

for all t ∈ R and x ∈ Aff K. We claim that f is positive. Given w ∈ D+\{0},<br />

write w = t(mg, g) + (x, −x) for some t ∈ R and x ∈ Aff K. Then tmg + x ≥ 0<br />

and tg − x ≥ 0 in Aff K. Adding these inequalities, we find that t(m + 1)g ≥ 0<br />

and hence that t > 0. In addition, 0 ≤ tg − x ≤ tg + tmg = t(m + 1)g, whence<br />

0 ≤ t −1 (m + 1) −1 (tg − x) ≤ g and so t −1 (m + 1) −1 p(tg − x) ≤ β. Then tp(g) −<br />

p(x) ≤ t(m + 1)β, and consequently −p(x) ≤ tmβ + t(β − p(g)) ≤ tmβ + tmǫ,<br />

whence f(w) = tm(β + ǫ) + p(x) ≥ 0. Thus f is positive as claimed. In particular<br />

f’s norm, relative to · ′ , is f(mg, g). Note that (mg, g) is an interior point of<br />

Aff K+ × Aff K+, so that the order unit norm · ′ on D is the restriction of the<br />

order unit norm on Aff K ⊕ Aff K defined by (mg, g), cf. Example 4.6 2). Extend<br />

f norm-preservingly to Aff K ⊕ Aff K by using the Hahn-Banach theorem, and<br />

conclude from Lemma 4.7 that the extension is positive on Aff K+ × Aff K+. We<br />

let ˜ f denote the extension, and define q : Aff K → R by q(x) = ˜ f(x, 0). Then q ≥ 0<br />

and q(x) = ˜ f(x, 0) ≥ ˜ f(x, 0) − ˜ f(0, x) = f(x, −x) = p(x), x ∈ Aff K+. Hence q ≥ p.<br />

Since mq(g) = ˜ f(mg, 0) ≤ ˜ f(mg, 0) + ˜ f(0, g) = f(mg, g) = m(β + ǫ), we see that<br />

q(g) ≤ β + ǫ, proving that<br />

inf{q(g) : q ∈ (Aff K) ∗ +, q ≥ p} ≤ β + ǫ.<br />

Since ǫ > 0 was arbitrary, the proof is complete. <br />

We can now derive the desired geometric description of the Choquet simplices.<br />

To get the must useful formulation we introduce an additional partial ordering


110 1. FUNDAMENTALS<br />

4) Every element of K is the barycenter of a unique Borel probability measure<br />

supported on the extreme boundary ∂eK.<br />

5) The upper envelope of every continuous convex function is affine.<br />

Conditions 1),2) and 3) are equivalent also when K is not metrizable.<br />

Proof. We prove that 1) ⇒ 3) ⇒ 5) ⇒ 4) ⇒ 3) ⇒ 2) ⇒ 1). The reader should<br />

check that the implications 1) ⇒ 3) ⇒ 2) ⇒ 1) make no use of the assumption that<br />

K is metrizable so that the 3 first conditions remain equivalent in the general case.<br />

The implication 1) ⇒ 3) follows from Lemma 4.23. 3) ⇒ 5) : Let f be a continuous<br />

convex function. We must prove that f is affine. To do this we identify K with<br />

S(Aff K), cf. Theorem 4.3. Extend f ∈ S(K) to (Aff K) ∗ + by setting<br />

f(tω) = tf(ω), t ≥ 0, ω ∈ S(Aff K).<br />

By using that f is convex on K it follows easily that f is subadditive on (Aff K) ∗ + ,<br />

i.e. that f(x + y) ≤ f(x) + f(y), x, y ∈ (Aff K) ∗ + . When we extend f to all of<br />

(Aff K) ∗ + in the same way, the concavity of f on K shows that f is superadditive on<br />

(Aff K) ∗ + , i.e. f(x + y) ≥ f(x) + f(y), x, y ∈ (Aff K)∗ + . It follows from Lemma 4.24<br />

that<br />

f(x) = sup{<br />

n<br />

f(yi) : yi ∈ (Aff K) ∗ + ,<br />

n<br />

yi = x}<br />

i=1<br />

i=1<br />

for all x ∈ (Aff K) ∗ + . Since Aff K∗ is a lattice it has the Riesz decomposition property,<br />

so when<br />

x + y = z1 + z2 + · · · + zn, x, y, z1, z2, · · · , zn ∈ (Aff K) ∗ +<br />

we may write zi = ai + bi, where 0 ≤ ai ≤ x, 0 ≤ bi ≤ y in Aff K∗ for all i and<br />

<br />

i ai = x, <br />

i bi = y. Hence<br />

<br />

f(zi) = <br />

f(ai + bi) ≤ <br />

f(ai) + <br />

f(bi) ≤ f(x) + f(y).<br />

i<br />

i<br />

i<br />

By taking the supremum over all such zi’s we get that f(x + y) ≤ f(x) + f(y),<br />

proving that f is additive on (Aff K) ∗ + . It follows that f is affine on K. (Note that<br />

we don’t know if f is continuous; only that it is upper semi-continuous.)<br />

5) ⇒ 4) : By using that the upper envelope of every continuous convex function<br />

is affine, we can now show that<br />

<br />

f(ω) = f dµ, µ ∈ M1(K), b(µ) = ω, (4.7)<br />

when f ∈ S(K). Since CR(K) is separable (because K is metrisable), we conclude<br />

from Exercise 4.12 that {g ∈ CR(K) : g ≥ f, −g ∈ S(K)} contains a dense sequence.<br />

There is therefore a sequence {hn} of concave continuous functions such that hn ≥ f<br />

for all n. Set<br />

gn = h1 ∧ h2 ∧ · · · ∧ hn .<br />

Then each gn is concave and majorises f. Furthermore, gn+1 ≤ gn for all n and<br />

limn→∞ gn = f pointwise on K. Therefore, by the Lebesgue theorem on monotone<br />

convergence, we find that<br />

<br />

f dµ = lim<br />

n<br />

gn dµ.<br />

i


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 111<br />

Let {µβ} be a net in M1(K) of finitely supported measures with barycenter ω and<br />

converging to µ. Such a net exists by Lemma 4.14. Since f is affine we have that<br />

<br />

f(ω) = f dµβ ≤ gn dµβ ≤ gn(ω)<br />

for all β and all n. By taking the limit over β we get that<br />

<br />

f(ω) ≤ gn dµ ≤ gn(ω)<br />

for all n. Now take the limit over n and conclude that<br />

<br />

f(ω) ≤ f dµ ≤ f(ω),<br />

proving (4.7). Now 4) follows easily : If µ1, µ2 ∈ M1(K) are both supported on ∂eK<br />

and have barycenter ω, it follows from Lemma 4.16 that fdµi = f dµi, f ∈<br />

CR(K), i = 1, 2. So (4.7) yields the conclusion that<br />

<br />

<br />

fdµ1 = f dµ1 = f(ω) = f dµ2 = f dµ2, f ∈ S(K).<br />

It follows now from Lemma 4.18 that µ1 = µ2.<br />

4) ⇒ 3) : Let l, t ∈ Aff K ∗ . By Lemma 4.9 we can write l = l1 − l2, t = t1 − t2<br />

where l1, l2, t1, t2 ∈ (Aff K) ∗ + . Note that r ∈ Aff K∗ will be an infimum of l and<br />

t if and only if r + l2 + t2 is an infimum of l + l2 + t2 and t + l2 + t2. So to<br />

show that an infimum exists for every pair l, t, it suffices to consider the case where<br />

l, t ∈ (Aff K) ∗ + . By assumption there are unique positive Borel measures µ, ν on<br />

K such that l(f) = f dµ and t(f) = f dν for all f ∈ Aff K and µ(K\∂eK) =<br />

ν(K\∂eK) = 0. Consider ν and µ as elements of CR(K) ∗ and let δ be the infimum<br />

of ν and µ in CR(K) ∗ , cf. Lemma 4.23. By the Riesz representation theorem, δ is<br />

given by a positive Borel measure on K; we denote this measure by δ also. Since<br />

f dδ ≤ min{ f dµ, f dν} for all f ∈ Aff K+, we see that δ|Aff K is a lower bound<br />

of l and t. To show that δ|Aff K is the greatest lower bound for l and t, let r ∈ Aff K∗ such that r ≤ t and r ≤ l. Write r = r1 − r2 where r1, r2 ∈ Aff K∗ +. Then r2(f) =<br />

f dχ, f ∈ Aff K, for some Borel measure χ supported on ∂eK. Set l1 = l +r2 and<br />

t1 = t + r2. Then 0 ≤ r + r2 ≤ l1 and r + r2 ≤ t1. Let µ1 and ν1 be the positive<br />

Borel measures supported on ∂eK such that l1(f) = f dµ1 and t1(f) = f dν1<br />

for all f ∈ Aff K. Then µ1 = µ + χ and ν1 = ν + χ. Let δ1 denote the infimum in<br />

CR(K) ∗ of the functionals CR(K) ∋ f ↦→ f dµ1 and CR(K) ∋ f ↦→ fdν1. Then<br />

δ1(f) = δ(f) + f dχ for all f ∈ CR(K). Since 0 ≤ r + r2 there is a positive Borel<br />

measure ǫ supported on ∂eK such that r(f) + r2(f) = f dǫ, f ∈ Aff K. Since<br />

l1 ≥ r + r2 and t1 ≥ r + r2 in Aff K∗ , there are positive Borel measures ǫl and ǫt,<br />

both supported on ∂eK, such that<br />

<br />

f dµ1 = f dǫ + f dǫl<br />

and <br />

<br />

f dν1 =<br />

<br />

f dǫ +<br />

f dǫt<br />

for all f ∈ Aff K. By the uniqueness assumption it follows that µ1 = ǫ + ǫl and<br />

ν1 = ǫ + ǫt. Therefore fdǫ ≤ f dµ1 and fdǫ ≤ f dν1 for all f ∈ CR(K)+, so


112 1. FUNDAMENTALS<br />

because δ1 is the greatest lower bound for CR(K) ∋ f ↦→ f dν1 and CR(K) ∋ f ↦→<br />

fdµ1, we conclude that <br />

f dǫ ≤ δ1(f), f ∈ CR(K)+.<br />

Restricted to f ∈ Aff K+ this shows that r + r2 ≤ δ1|Aff K = δ|Aff K + r2, i.e.<br />

r ≤ δ|Aff K.<br />

3) ⇒ 2) : We prove that the strict ordering has the Riesz decomposition property.<br />

Given f1, f2 ∈ Aff K+ such that fi >> 0, i = 1, 2, set<br />

Fi = {f ∈ Aff K : 0


and h < δǫ. Then set<br />

4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 113<br />

hi = (1 + 2δ) −1 (gi + h + δǫ), i = 1, 2,<br />

and note that h1 + h2 = f. Since h < δǫ, we have that<br />

and hence 0


114 1. FUNDAMENTALS<br />

Lemma 4.31. Let {xn} and {yn} be sequences in A such that<br />

<br />

=<br />

n<br />

xnx ∗ n<br />

m<br />

y ∗ m ym .<br />

It follows that there is a sequence {znm : n, m ∈ N} such that<br />

<br />

z ∗ nizni = x ∗ nxn, <br />

for all n, m.<br />

i<br />

i<br />

zimz ∗ im = ymy ∗ m<br />

Proof. Set x = <br />

n xnx∗ 1<br />

n . Since x is positive, k<br />

k ∈ N. For each m, n, k ∈ N, consider<br />

ym( 1<br />

+ x)−1 2xn .<br />

k<br />

+ x is invertible in  for all<br />

Let ǫ > 0 and set ak,l = ( 1 + x)−12<br />

− ( k 1 + x)−1 2. For all k, l ∈ N we have that<br />

l<br />

ym( 1<br />

k<br />

+ x)−1 2xn − ym( 1<br />

l<br />

+ x)−12xn<br />

2 = ymak,lxn 2<br />

= x ∗ nak,ly ∗ mymak,lxn ≤ x ∗ nak,lxak,lxn<br />

since y ∗ mym ≤ x so that x ∗ nak,ly ∗ mymak,lxn ≤ x ∗ nak,lxak,lxn. But<br />

x ∗ nak,lxak,lxn = x 1<br />

2ak,lxnx ∗ 1<br />

nak,lx 2 ≤ x 1<br />

2ak,lxak,lx 1<br />

2 <br />

for a similar reason and hence<br />

ym( 1<br />

+ x)−12xn<br />

− ym(<br />

k 1<br />

+ x)−12xn<br />

l 2 ≤ x 2 a 2 k,l since x and ak,l commute. Since limk→∞ t2<br />

1<br />

k<br />

spectral theory that limk→∞ x2 ( 1<br />

k + x)−1 = x in A. By taking square-roots we see<br />

that limk→∞ x( 1<br />

+ x)−1 2 = x k 1<br />

2 in A. In particular, this shows that<br />

x 2 a 2 k,l = x2 ( 1<br />

< ǫ<br />

k + x)−1 + ( 1<br />

l + x)−1 − 2( 1<br />

k<br />

for all sufficiently large k, l. We conclude that {ym( 1<br />

k<br />

all n, m and we set znm = limk→∞ ym( 1<br />

k<br />

Note that<br />

+t = t, uniformly on [0, ∞[, we get from<br />

+ x)−12xn.<br />

+ x)−1 2( 1<br />

l<br />

<br />

+ x)−1 2 <br />

+ x)−1 2xn} is Cauchy in A for<br />

ym( 1<br />

k + x)−1xy ∗ m − ymy ∗ m = ym(1 − ( 1<br />

k + x)−1x)y ∗ m =<br />

(1 − ( 1<br />

k + x)−1x) 1<br />

2y ∗ mym(1 − ( 1<br />

k + x)−1x) 1<br />

2 ≤<br />

x − ( 1<br />

k + x)−1x 2 <br />

for all m and that limk→∞ x − ( 1<br />

k + x)−1 x 2 = 0. Fix now ǫ > 0 and m ∈ N.<br />

Choose N ∈ N so large that 0 ≤ x − N<br />

n=1 xnx ∗ n ≤ ǫ1, and take K ∈ N such that


ymy ∗ m<br />

4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 115<br />

1 − ym( k + x)−1xy∗ m ≤ ǫ1 for all k ≥ K. Then<br />

0 ≤ ymy ∗ m −<br />

N<br />

i=1<br />

ym( 1<br />

+ x)−12xix<br />

k ∗ i (1 + x)−1 2y<br />

k ∗ m<br />

≤ ymy ∗ m − ym( 1<br />

k + x)−1 xy ∗ m + ǫym( 1<br />

k + x)−1 y ∗ m ≤ ǫ1 + ǫym( 1<br />

k + x)−1 y ∗ m<br />

for all k ≥ K. Note that<br />

ym( 1<br />

k + x)−1y ∗ m = ( 1<br />

+ x)−12y<br />

k ∗ mym( 1<br />

+ x)−12<br />

<br />

k<br />

≤ ( 1<br />

+ x)−12x(<br />

k 1<br />

+ x)−12<br />

= <br />

k x<br />

so that ǫ1 + ǫym( 1<br />

k + x)−1 y ∗ m ≤ 2ǫ1. It follows that<br />

0 ≤ ymy ∗ m −<br />

N<br />

i=1<br />

<br />

= lim ymy<br />

k→∞<br />

∗ m −<br />

zimz ∗ im<br />

N<br />

i=1<br />

1<br />

k<br />

+ x ≤ 1 ,<br />

ym( 1<br />

+ x)−12xix<br />

k ∗ i( 1<br />

+ x)−1 2y<br />

k ∗ <br />

m ≤ 2ǫ1 ,<br />

proving that ymy∗ m − N i zimz∗ im. By<br />

exchanging m with n, ym with x∗ n and xn with ym in the previous argument we get<br />

that x∗ nxn = <br />

i z∗ nizni. <br />

i=1 zimz ∗ im ≤ 2ǫ. This shows that ymy ∗ m = <br />

When a, b are positive elements in A we write a ≈ b when there is a sequence<br />

{xn} in A such that a = ∞ n=1 x∗nxn and b = ∞ n=1 xnx∗ n. ≈ is an equivalence relation<br />

on A+ thanks to Lemma 4.31. Indeed, x = x 1<br />

2x 1<br />

2 = x for all x ∈ A+, showing that<br />

x ≈ x. If {xn} is a sequence establishing a ≈ b then {x∗ n } gives that b ≈ a. To<br />

prove transitivity we use Lemma 4.31. If namely a = <br />

n x∗nxn, b = <br />

n xnx∗ <br />

n, b =<br />

<br />

i z∗ ni zni = x ∗ n xn and <br />

i zimz ∗ im = ymy ∗ m<br />

m y∗ mym and c = <br />

m ymy ∗ m, then Lemma 4.31 gives us a sequence {znm} such that<br />

a while <br />

n,m znmz ∗ nm<br />

= <br />

m ymy ∗ m<br />

for all n. Then <br />

n,m z∗ nm znm = <br />

n x∗ n xn =<br />

= c, proving that a ≈ c. We next extend this<br />

equivalence relation to the selfadjoint part Asa of A by setting a ∼ b when there is<br />

a positive element p ∈ A+ such that a + p ≈ b + p. Again we argue that ∼ is an<br />

equivalence relation on Asa. If a = a∗ we know that a = a+ − a−, a+, a− ∈ A+, so<br />

a + a− = (a+) 1<br />

2(a+) 1<br />

2 = a + a−, showing that a ∼ a. That ∼ is symmetric on Asa is<br />

obvious. If a ∼ b, b ∼ c in Asa, then there are positive elements p, q ∈ A+ such that<br />

a + p ≈ b + p and b + q ≈ c + q in A+. Hence a + p + q ≈ b + p + q ≈ c + p + q in<br />

A+, showing that a ∼ c in Asa. For a ∈ Asa we denote the equivalence class under<br />

∼ containing a by [a].<br />

We observe that a ∼ b if and only if there are positive elements a1, a2, b1, b2 ∈ A+<br />

such that a = a1 − a2, b = b1 − b2 and a1 + b2 ≈ b1 + a2. Indeed, when a ∼ b and<br />

p ∈ A+ such that a + p ≈ b + p, then we write a = a+ − a−, b = b+ − b−, where<br />

a+, a−, b+, b− ∈ A+ and note that a++2p+b− = a+2p+b−+a− ≈ b+2p+b−+a− =<br />

b+ + 2p + a−, a = a+ + p − (a− + p), and b = b+ + p − (b− + p). Conversely, when<br />

a = a1−a2, b = b1−b2, a1, a2, b1, b2 ∈ A+ and a1+b2 ≈ b1+a2, then p = a2+b2 ∈ A+<br />

and a + p = a1 + b2 ≈ b1 + a2 = b + p.


116 1. FUNDAMENTALS<br />

Lemma 4.32. Asa/ ∼ is a real vector space with the operations<br />

t[a] + [b] = [ta + b], t ∈ R, a, b ∈ Asa.<br />

Proof. It suffices to check that the addition and scalar multiplication is welldefined<br />

on Asa/ ∼. We must show that when a, b, c ∈ Asa, t ∈ R and a ∼ c, then<br />

ta ∼ tc and a+b ∼ c+b. If t ≥ 0 we write a = a1−a2, b = b1−b2, c = c1−c2 such that<br />

a1+c2 ≈ c1+a2. Then ta1+tc2 ≈ tc1+ta2, showing that ta ∼ tc. If instead t ≤ 0, we<br />

observe that −ta1 − tc2 ≈ −tc1 − ta2 while ta = −ta2 − (−t)a1, tc = −tc2 − (−t)c1,<br />

which shows that ta ∼ tc. We leave the reader to show that a + b ∼ c + b. <br />

Set A = Asa/ ∼. We go on to show that A is a Banach space.<br />

Lemma 4.33. Let x, y, a, b ∈ A+ such that a ≈ b and x + a ≈ y + b. Then<br />

b for all n ∈ N.<br />

x + 1 1 a ≈ y + n n<br />

Proof. Note that x + 1 1 a ≈ y + b ⇔ nx + a ≈ ny + b. We prove the last<br />

n n<br />

equivalence by induction in n. By assumption the equivalence holds in the case<br />

n = 1. Assume that it has been established for n. Then<br />

(n + 1)x + a ≈ x + ny + b ≈ x + ny + a ≈ ny + y + b = (n + 1)y + b.<br />

Lemma 4.34. Set A0 = {x−y : x, y ∈ A+, x ≈ y}. Then A0 is a closed subspace<br />

of Asa.<br />

Proof. It is straightforward to check that A0 is a subspace of Asa. Let z ∈ A0<br />

and ǫ > 0. We show that<br />

z = x − y for some x, y ∈ A+ such that x ≈ y and x + y ≤ z + ǫ . (4.8)<br />

We have that z = a − b with a, b ∈ A+, a ≈ b. Since z is selfadjoint we can write<br />

z = z+ − z− where z+, z− ∈ A+, z+z− = 0 and max{z+, z−} = z. Then<br />

z+ +b = z− +a, whence z+ + 1<br />

nb ≈ z− + 1a<br />

for all n ∈ N by Lemma 4.33. Let n ∈ N<br />

n<br />

be so large that 2<br />

na + b < ǫ and set x = z+ + 1<br />

n (a + b), y = z− + 1(a<br />

+ b). Then<br />

n<br />

1 b) + b = y. Furthermore,<br />

z = x − y and x = (z+ + 1<br />

n<br />

na ≈ (z− + 1 1 a) + n n<br />

x+y = z+ +z− + 2<br />

n (a+b) ≤ z+ +z−+ 2<br />

n (a+b) ≤ z++z−+ǫ = z+ǫ .<br />

We then use (4.8) to show that A0 is closed. Let {zn} be a sequence in A0 such that<br />

limn→∞ zn = z in Asa. To show that z ∈ A0 we pass to a subsequence {zni } such<br />

that zni − zni+1 < 2−i for all i. By applying (4.8) to zni+1 − zni we get elements<br />

xi, yi ∈ A+ such that zni+1 − zni = xi − yi, xi ≈ yi and xi + yi ≤ 2 −i . Then<br />

x = ∞ i=1 xi and y = ∞ i=1 yi exist in A+. Since xi ≈ yi for all i, we see that x ≈ y.<br />

Furthermore<br />

∞<br />

∞<br />

z − zn1 = − zni ) = (xi − yi) = x − y ,<br />

i=1<br />

(zni+1<br />

proving that z = x − y + zn1 ∈ A0. <br />

Since x ∼ y in Asa if and only if x − y ∈ A0, we have that A = Asa/A0. (Note<br />

that this description of ∼ gives an alternative approach to Lemma 4.32.) By Lemma<br />

4.34 A is a Banach space in the quotient norm<br />

i=1<br />

[x] = inf{x − z : z ∈ A0} .


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 117<br />

To proceed further with the investigation of the structure of A, in the unital case,<br />

we need the following decomposition theorem for functionals on a <strong>C∗</strong>-algebra. Recall<br />

that a continuous functional ω on a <strong>C∗</strong>-algebra A is selfadjoint when ω(a) = ω(a∗ ),<br />

or, alternatively, ω(Asa) ⊆ R. The formula<br />

ω(a) = 1<br />

2 (ω(a) + ω(a∗ )) + i( 1<br />

2i (ω(a) − ω(a∗ )))<br />

shows that an arbitrary element ω ∈ A∗ is a linear combination of two selfadjoint<br />

continuous functionals.<br />

Theorem 4.35. (Hahn-Jordan decomposition) Let A be a unital C ∗ -algebra and<br />

ω ∈ A ∗ be a selfadjoint functional. There are unique positive linear functionals<br />

ω+, ω− ∈ A ∗ such that ω = ω+ − ω− and ω = ω+ + ω−.<br />

Proof. The theorem is trivial when ω = 0 so we consider only the case where<br />

ω = 0. It follows from Lemma 4.9 that there are states µ1, µ2 ∈ S(A) and a number<br />

t ∈ [0, 1] such that<br />

ω −1 ω = tµ1 − (1 − t)µ2 .<br />

Set ω+ = ωtµ1 and ω− = ω(1 − t)µ2. This give the existence part. To prove<br />

the uniqueness assume that ω = µ+ − µ− where µ+, µ− are positive functionals on<br />

A with µ+ + µ− = ω. We must show that ω+ = µ+ and µ− = ω−. Let ǫ > 0.<br />

Since ω is selfadjoint there is a selfadjoint element a in the unit ball of A such that<br />

ω(a) > ω − 1<br />

(1 −a) and note that 0 ≤ b ≤ 1. We have that<br />

2ǫ2 (check !). Set b = 1<br />

2<br />

ω+(1) + ω−(1) = ω+ + ω− = ω < ω(a) + 1<br />

2 ǫ2 = ω+(a) − ω−(a) + 1<br />

2 ǫ2 ,<br />

showing that ω+(1 − a) + ω−(1 + a) < 1<br />

2 ǫ2 or<br />

ω+(b) + ω−(1 − b) < 1<br />

4 ǫ2 .<br />

It follows that 0 ≤ ω+(b) < 1<br />

inequality, Lemma 1.78, we have that<br />

and<br />

4ǫ2 and 0 ≤ ω−(1 − b) < 1<br />

4ǫ2 . By using the Schwarz<br />

|ω+(bx)| 2 = |ω+(b 1<br />

2b 1<br />

2x)| 2 ≤ ω+(b)ω+(x ∗ bx)<br />

≤ ω+(b)ω+x 2 < 1<br />

4 ǫ2ω+x 2<br />

|ω−((1 − b)x)| 2 = |ω((1 − b) 1<br />

2(1 − b) 1<br />

2x)| 2 ≤ ω−(1 − b)ω−(x ∗ (1 − b)x)<br />

< 1<br />

4 ǫ2ω−x 2<br />

for all x ∈ A. Hence |ω+(bx)| ≤ 1<br />

2ǫω1 2 x and |ω−((1 − b)x)| ≤ 1<br />

2ǫω1 2 x<br />

for all x ∈ A. The same argument applies to µ+ and µ−, so we find that also<br />

|µ+(bx)| ≤ 1<br />

2ǫω1 2 x and |µ−((1 − b)x)| ≤ 1<br />

2ǫω1 2 x for all x ∈ A. But since<br />

ω+ − ω− = µ+ − µ−, we find that<br />

ω+(x) − µ+(x) = ω+(bx) − µ+(bx) + ω−((1 − b)x) − µ−((1 − b)x)<br />

and so |ω+(x) − µ+(x)| ≤ 2ǫω 1<br />

2 x for all x ∈ A. Since ǫ > 0 was arbitrary, we<br />

see that ω+ = µ+ and hence also ω− = µ−. <br />

In the following A will denote a unital C ∗ -algebra.


118 1. FUNDAMENTALS<br />

Definition 4.36. An element ω ∈ A ∗ is tracial when ω(xy) = ω(yx), x, y ∈ A.<br />

A tracial state on A, i.e. a tracial element of S(A), will be called a trace state. The<br />

set of trace states of A will be denoted by T(A).<br />

Note that T(A) is a convex weak ∗ -compact subset of A ∗ .<br />

Lemma 4.37. Let ω ∈ A ∗ . Then the following conditions are equivalent.<br />

1) ω is tracial.<br />

2) ω(xx ∗ ) = ω(x ∗ x), x ∈ A.<br />

3) ω(uxu ∗ ) = ω(x) for all x ∈ A and all unitaries u in A.<br />

Proof. 1) ⇒ 2) is trivial. 2) ⇒ 3) : When x ≥ 0 we find that ω(uxu∗ ) =<br />

ω(ux 1<br />

2x 1<br />

2u∗ ) = ω(x 1<br />

2u∗ux 1<br />

2) = ω(x). Since an arbitrary element x of A is a linear<br />

combination of 4 positive elements, 3) follows. 3) ⇒ 1) : Assume first that y = u<br />

is a unitary. Then ω(xy) = ω(xu) = ω(u∗ (ux)u) = ω(ux) = ω(yx), showing that 1)<br />

holds in this case. The general case follows from the fact that an arbitrary element<br />

y is a linear combination of 4 unitaries. See Exercise 4.38 below. <br />

Exercise 4.38. Let y ∈ A. Show that there are 4 unitaries u1, u2, u3, u4 in A<br />

and scalars λ1, λ2, λ3, λ4 ∈ C such that y = λ1u1 + λ2u2 + λ3u3 + λ4u4.<br />

(Hint : After a few frustrations the reader will realize that it suffices to show that<br />

a real-valued continuous function on a compact Hausdorff space, taking values in<br />

[−1, 1], is the mean of two continuous functions taking values on T. Then remember<br />

that 2t = t + i √ 1 − t 2 + t − i √ 1 − t 2 , t ∈ [−1, 1]. )<br />

We can now easily obtain the Hahn-Jordan decomposition for self- adjoint trace<br />

functionals.<br />

Theorem 4.39. Let ω ∈ A ∗ be a selfadjoint tracial functional. There are then<br />

unique positive tracial functionals ω+, ω− such that ω = ω+ −ω− and ω = ω++<br />

ω−.<br />

Proof. Consider the Hahn-Jordan decomposition ω = ω+ − ω− given by Theorem<br />

4.35. Let u ∈ A be a unitary and set µ+(·) = ω+(u · u ∗ ), µ−(·) = ω−(u · u ∗ ).<br />

Since ω is a tracial functional, ω = µ+ − µ−, cf. Lemma 4.37. Furthermore, since<br />

uxu ∗ = x for all x ∈ A, we have that µ± = ω±. The uniqueness of the<br />

Hahn-Jordan decomposition implies that µ± = ω±, i.e. ω±(uxu ∗ ) = ω±(x), x ∈ A.<br />

Since this conclusion holds for all unitaries u in A, we conclude from Lemma 4.37<br />

that ω± are both tracial. The uniqueness statement follows immediately from the<br />

uniqueness of the Hahn-Jordan decomposition, Theorem 4.35.<br />

<br />

Now we define a closed convex cone A+ in A by<br />

A+ = {[a] : a ∈ A+} ,<br />

i.e. if we let q : Asa → A be the quotient map, A+ = q(A+). Let u = q(1).<br />

Lemma 4.40. A is an order unit space with positive cone A+ = q(A+) and order<br />

unit u = q(1).<br />

Proof. We must show that {x ∈ A : x ≤ 1} = {x ∈ A : −u ≤ x ≤ u}.<br />

Assume first that x ≤ 1. Then x = q(a) for some a ∈ Asa. For every z ∈ A0 we<br />

have that −a − z1 ≤ a − z ≤ a − z1. Hence −a − zu ≤ q(a) = x ≤ a − zu


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 119<br />

in A. Since x = inf{a − z : z ∈ A0} ≤ 1 and since A+ is closed, it follows that<br />

−u ≤ x ≤ u.<br />

Conversely, assume that −u ≤ x ≤ u. To prove that x ≤ 1, we use the<br />

Hahn-Banach theorem to find l ∈ A ∗ such that l = 1 and l(x) = x. Set<br />

ω = l ◦ q ∈ (Asa) ∗ and extend ω to a selfadjoint functional on A by ω(a + ib) =<br />

ω(a) + iω(b), a, b ∈ Asa. Then ω is a selfadjoint trace functional on A and we write<br />

ω = ω+ −ω− where ω± are positive trace functionals on A and ω = ω++ω−.<br />

Since ω± annihilate A0, we can define l± ∈ A ∗ by l± ◦ q = ω±. Note that since A<br />

has the quotient norm, l± = ω± and l = l ◦ q = ω. Hence l = l+ − l−<br />

and l = l+ + l−. Since −u ≤ x ≤ u we have that<br />

−l±(u) ≤ l±(x) ≤ l±(u) .<br />

Since l±(u) = ω±(1) = ω±, we see that |l±(x)| ≤ l± and hence<br />

x = l(x) = l+(x) − l−(x) ≤ l+ + l− .<br />

Since l+ + l− = l = 1 we have that x = l(x) ≤ 1. <br />

The next aim is to show that the order unit space A is a copy of Aff T(A). Define<br />

Φ : A → Aff T(A) by Φ(q(a))(ω) = ω(a), ω ∈ T(A), a ∈ Asa.<br />

Lemma 4.41. Φ is a linear isometry such that Φ(A+) = Aff T(A)+ and Φ(u) = 1.<br />

Proof. Clearly, Φ is a linear map taking u to the constant function 1 and A+<br />

into Aff T(A)+. Let Λ : A → Aff S(A) be the isomorphism of Theorem 4.10, i.e.<br />

Λ(x)(l) = l(x), x ∈ A, l ∈ S(A). It suffices to show that Φ ◦ Λ −1 : Aff S(A) →<br />

Aff T(A) is an isometry taking Aff S(A)+ onto Aff T(A)+. To this end, define ψ :<br />

T(A) → S(A) by ψ(ω)(q(a)) = ω(a), a ∈ Asa, ω ∈ T(A). Then Φ ◦ Λ −1 (f)(ω) =<br />

f(ψ(ω)), f ∈ Aff S(A), ω ∈ T(A). It therefore suffices to show that ψ is an affine<br />

homeomorphism. To this end define ϕ : S(A) → T(A) by ϕ(l)(a) = l(q(a)), a ∈<br />

Asa, l ∈ S(A). Then ψ and ϕ are both affine and are in fact inverses of each other.<br />

Indeed, ϕ ◦ ψ(ω)(a) = ψ(ω)(q(a)) = ω(a), ω ∈ T(A), a ∈ Asa, and ψ ◦ ϕ(l)(q(a)) =<br />

ϕ(l)(a) = l(q(a)), l ∈ S(A), a ∈ Asa.<br />

<br />

Corollary 4.42. Let A be a unital C ∗ -algebra. For each f ∈ Aff T(A) there is<br />

a self-adjoint element a = a ∗ ∈ A such that a = f. (Recall that a(ω) = ω(a), ω ∈<br />

T(A).)<br />

The partial ordering of A∗ sa gives rise to a partial ordering of the selfadjoint trace<br />

functionals on A in the obvious way; µ1 ≤ µ2 when µ1(a) ≤ µ2(a) for all a ∈ A+.<br />

Lemma 4.43. Let ω1, ω2 ∈ A ∗ be selfadjoint trace functionals on A. There is then<br />

a selfadjoint trace functional µ ∈ A ∗ such that µ ≥ ωi, i = 1, 2, and when ν ∈ A ∗ is<br />

any selfadjoint trace functional on A such that ν ≥ ωi, i = 1, 2, then µ ≤ ν.<br />

In other words, the supremum of ω1 and ω2 exists in the space of selfadjoint trace<br />

functionals.<br />

Proof. By Theorem 4.39 we may write ωi = ω +<br />

i<br />

− ω−<br />

i<br />

where ω±<br />

i<br />

are positive<br />

trace functionals, i = 1, 2. So by considering µ1 = ω1+ω − 1 +ω− 2 and µ2 = ω2+ω − 1 +ω− 2<br />

in place of ω1 and ω2, we can assume that ωi ≥ 0, i = 1, 2. We first define µ on A+<br />

as<br />

µ(a) = sup{ω1(x) + ω2(y) : x, y ∈ A+, a ≈ x + y} .


120 1. FUNDAMENTALS<br />

Let s, t ∈ A+. When x, y, x1, y1 ∈ A+ such that s ≈ x + y, t ≈ x1 + y1, we find that<br />

ω1(x) + ω2(y) + ω1(x1) + ω2(y1) = ω1(x + x1) + ω2(y + y1) ≤ µ(s + t)<br />

since s + t ≈ (x + x1) + (y + y1). Thus µ(s) + µ(t) ≤ µ(s + t). On the other hand,<br />

consider x, y ∈ A+ such that s + t ≈ x + y. By definition this means that there is a<br />

sequence {xn} in A such that ∞ n=1 x∗n xn = s + t, ∞ n=1 xnx∗ n = x + y. But then<br />

s 1<br />

2s 1<br />

2 + t 1<br />

2t 1<br />

∞<br />

2 =<br />

n=1<br />

x ∗ n xn<br />

so<br />

<br />

Lemma 4.31 applies to give us two sequences {vn}, {wn} in A such that s =<br />

n v∗ nvn, t = <br />

<br />

= x+y. Now we apply Lemma<br />

n w∗ nwn and vkv∗ k +wkw∗ k = xkx∗ k for all k ∈ N. Set s′ = <br />

n vnv∗ n , t′ =<br />

n wnw∗ n . Then s ≈ s′ , t ≈ t ′ and s ′ +t ′ = <br />

k xkx∗ k<br />

4.31 again to get elements zij ∈ A, i, j ∈ {1, 2}, such that s ′ = z∗ 11z11 + z∗ 12z12, t ′ =<br />

z∗ 21z21 + z∗ 22z22, x = z11z∗ 11 + z21z∗ 21 and y = z12z∗ 12 + z22z∗ 22. Set as = z11z∗ 11, bs =<br />

z12z∗ 12 , at = z21z∗ 21 and bt = z22z∗ 22 . Then s ≈ s′ ≈ as + bs, t ≈ t ′ ≈ at + bt, and<br />

x = as + at, y = bs + bt. Therefore<br />

ω1(x) + ω2(y) = ω1(as) + ω2(bs) + ω1(at) + ω2(bt) ≤ µ(s) + µ(t) .<br />

By taking the supremeum over all x, y ∈ A+ with s + t ≈ x + y, we find that<br />

µ(s + t) ≤ µ(s) + µ(t). It follows that µ(s + t) = µ(s) + µ(t), i.e. µ is additive on<br />

A+. We can then extend µ to a linear functional on Asa by<br />

µ(a − b) = µ(a) − µ(b), a, b ∈ A+,<br />

and then to a selfadjoint functional on A by µ(a+ib) = µ(a)+iµ(b), a, b ∈ Asa. When<br />

0 ≤ b ≤ 1 we get immediately from the definition of µ that 0 ≤ µ(b) ≤ ω1 + ω2.<br />

(1+a) ≤ 1 and<br />

When −1 ≤ a ≤ 1, we have that a = 1<br />

2<br />

(1+a) − 1<br />

2<br />

(1 −a) where 0 ≤ 1<br />

2<br />

0 ≤ 1<br />

2 (1 − a) ≤ 1, so we conclude that −(ω1 + ω2) ≤ µ(a) ≤ ω1 + ω2. It<br />

follows that µ is continuous, i.e. that µ ∈ A ∗ . By definition µ is a trace functional<br />

since ω1 and ω2 are, and it is clear that µ ≥ ωi, i = 1, 2. If ν ∈ A ∗ is a trace<br />

functional such that ν ≥ ωi, i = 1, 2, then<br />

ω1(x) + ω2(y) ≤ ν(x + y) = ν(a)<br />

for all a, x, y ∈ A+ such that a ≈ x + y. Hence µ(a) ≤ ν(a). <br />

Theorem 4.44. Let A be a unital C ∗ -algebra. Then T(A) is a Choquet simplex.<br />

Proof. By Theorem 4.26 it suffices to show that the ordered dual Aff T(A) ∗ is<br />

a lattice. By Lemma 4.41 this is equivalent to A ∗ being a lattice. So let µ1, µ2 ∈ A ∗ .<br />

The complexifications of the functionals ωi = µi ◦ q, i = 1, 2, are selfadjoint trace<br />

functionals on A. Let µ be the supremum of ω1 and ω2 among the selfadjoint trace<br />

functionals on A, cf. Lemma 4.43. We can then define µ ′ ∈ A ∗ by µ ′ (q(x)) =<br />

µ(x), x ∈ Asa. It is straigthforward to check that µ ′ is the supremum of µ1 and µ2<br />

in A ∗ . <br />

4.4. Examples of trace spaces.<br />

Lemma 4.45. The only trace state on the <strong>C∗</strong>- algebra Mn of complex n × n<br />

matrices is given by<br />

τ (aij) = 1<br />

n<br />

aii , (aij) ∈ Mn .<br />

n<br />

i=1


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 121<br />

Proof. We first check that τ is a trace state. When a = (aij) ∈ Mn we have<br />

that aa ∗ = ( n<br />

k=1 aikajk) so that<br />

τ(aa ∗ ) = 1<br />

n<br />

<br />

|aik| 2 ≥ 0 .<br />

i,k<br />

This shows that τ is positive. It is clear that τ(1) = 1 so τ is a state on Mn. Let<br />

{eij} be the standard system of matrix units in Mn. Then τ(eijekl) = δjkτ(eil) =<br />

1<br />

n δjkδil = τ(ekleij). Since arbitrary elements x, y ∈ Mn are linear combinations of<br />

the matrix units it follows by linearity that τ(xy) = τ(yx), i.e. τ is a trace state.<br />

Let ω be a trace state on Mn. Then<br />

ω(ekl) = ω(ekkeklell) = ω(ellekkekl) = δklω(ekk) .<br />

Since ω(e11) = ω(e1kek1) = ω(ek1e1k) = ω(ekk), k = 1, 2, · · · , n, and 1 = ω(1) =<br />

ω( n i=1 eii) = n i=1 ω(eii) = nω(e11), it follows that ω = τ. <br />

From Lemma 4.45 it is easy to give the following description of the tracial state<br />

space T(F) of the finite dimensional C ∗ -algebra F = Mn1 ⊕Mn2 ⊕· · ·⊕Mnm. Denote<br />

the trace state of Mni by τi, i = 1, 2, · · · , m. For each vector (t1, t2, · · · , tm) ∈ R m<br />

we define a linear functional τ(t1,t2,···,tm) : F → C by<br />

m<br />

τ(t1,t2,···,tm)(a1, a2, · · · , am) = tiτi(ai) .<br />

Lemma 4.46. The map (t1, t2, · · · , tm) ↦→ τ(t1,t2,··· ,tm) defines an affine homeomorphism<br />

from ∆m onto T(F).<br />

Proof. We leave the reader to check that the map takes ∆m affinely into T(F).<br />

It is injective because<br />

i=1<br />

τ(t1,t2,···,tm)(0, 0, · · · , 0, 1, 0, 0, · · · , 0)<br />

= ti .<br />

1 at the i’th entry<br />

To prove the surjectivity, let ω ∈ T(F) and let pi = (0, 0, · · · , 0, 1, 0, 0, · · · , 0).<br />

Then<br />

1 at the i’th entry<br />

ti = ω(pi) ∈ [0, 1] and m i=1 ti = ω( <br />

i pi) = ω(1) = 1, i.e. (t1, t2, · · · , tm) ∈ ∆m.<br />

We leave the reader to use the uniqueness of the trace state on Mni to confirm that<br />

ω = τ(t1,t2,···,tm). <br />

We now consider C ∗ -<strong>algebras</strong> of the form C(X, A) where X is a compact Hausdorff<br />

space and A is a unital C ∗ -algebra, cf. Example 1.20. To work with such<br />

<strong>algebras</strong> it is convenient to introduce the following notation. When f : X → C is a<br />

continuous function and a ∈ A we define an element f ⊗ a ∈ C(X, A) by<br />

We shall need the following lemma.<br />

f ⊗ a(x) = f(x)a, x ∈ X.<br />

Lemma 4.47. For every g ∈ C(X, A) and every ǫ > 0 there are finite sets<br />

{f1, f2, · · · , fn} ⊆ C(X) and {a1, a2, · · · , an} ⊆ A such that g − n<br />

i=1 fi ⊗ ai ≤ ǫ.<br />

Proof. By compactness of X and the continuity of g we can find a finite open<br />

cover {Ui : i = 1, 2, · · · , n} of X and points xi ∈ Ui such that g(xi) − g(x) < ǫ<br />

for all x ∈ Ui, i = 1, 2, · · · , n. Set ai = g(xi) for all i and let {hi} be a partition of


122 1. FUNDAMENTALS<br />

unity in C(X) subordinate to {Ui : i = 1, 2, · · · , n}, i.e. hi ≥ 0 and supp(hi) ⊆ Ui<br />

for all i and n<br />

i=1 hi = 1. Then<br />

g(x) −<br />

n<br />

hi ⊗ ai(x) = <br />

i=1<br />

n<br />

hi(x)g(x) − g(xi) ≤<br />

i=1<br />

n<br />

hi(x)g(x) −<br />

i=1<br />

n<br />

hi(x)ǫ = ǫ.<br />

i=1<br />

n<br />

hi(x)g(xi) ≤<br />

Lemma 4.48. Assume that A is unital and has a unique trace state τ. Then<br />

T(C(X, A)) is affinely homeomorphic to M1(X). The trace state τµ ∈ T(C(X, A))<br />

corresponding to µ ∈ M1(X) is given by<br />

<br />

τµ(g) = τ(g(x)) dµ(x), g ∈ C(X, A).<br />

X<br />

Proof. It is straightforward to check that the map µ ↦→ τµ is a continuous<br />

affine map from M1(X) into T(C(X, A)). Since τµ(f ⊗ 1) = <br />

f dµ, f ∈ C(X),<br />

X<br />

we see immediately that the map is injective. To prove that it is also surjective, let<br />

ρ ∈ T(C(X, A)). The map C(X) ∋ f ↦→ ρ(f ⊗ 1) is a state on C(X), so it is given<br />

by a regular Borel probability measure µ ∈ M1(X) through the formula<br />

<br />

f dµ = ρ(f ⊗ 1), f ∈ C(X).<br />

X<br />

We assert that τµ = ρ. To check this it suffices by Lemma 4.47 to check that the<br />

continuous functionals agree on elements of the form f ⊗a where f ∈ C(X), a ∈ A.<br />

In fact, since an arbitrary f is a linear combination of 4 positive continuous functions<br />

it suffices to check the agreement when f ≥ 0. We consider first the case where<br />

ρ(f ⊗ 1) = 0. Then f dµ = 0 and f must be zero almost everywhere with respect<br />

to µ. Hence τµ(f ⊗ a) = <br />

τ(a)f(x) dµ(x) = 0. On the other hand the Schwarz<br />

X<br />

inequality gives that<br />

|ρ(f ⊗ a)| 2 = |ρ((f ⊗ 1)(1 ⊗ a))| 2 ≤ ρ(f 2 ⊗ 1)ρ(1 ⊗ a ∗ a) = ρ(1 ⊗ a ∗ <br />

a) f 2 dµ = 0 ,<br />

i.e. ρ(f ⊗ a) = 0. Thus the two functionals agree on f ⊗ a in this case. Assume<br />

next that ρ(f ⊗ 1) = 0. Then χ(a) = ρ(f ⊗ 1) −1 ρ(f ⊗ a), a ∈ A, defines a state on<br />

A. Let x, y ∈ A. Since (1 ⊗ y)(f ⊗ 1) = (f ⊗ 1)(1 ⊗ y) = f ⊗ y and ρ is a trace on<br />

C(X, A), we find that<br />

ρ(f ⊗ 1)χ(xy) = ρ(f ⊗ xy) = ρ((f ⊗ 1)(1 ⊗ x)(1 ⊗ y)) = ρ((1 ⊗ y)(f ⊗ 1)(1 ⊗ x))<br />

= ρ((f ⊗ 1)(1 ⊗ y)(1 ⊗ x)) = ρ(f ⊗ yx) = ρ(f ⊗ 1)χ(yx) ,<br />

proving that χ is a trace state on A. By the uniquenes of the trace state of A, we<br />

conclude that χ = τ. It follows that<br />

<br />

ρ(f ⊗ a) = ρ(f ⊗ 1)χ(a) = τ(a) f dµ = τ(f ⊗ a(x)) dµ(x) = τµ(f ⊗ a) .<br />

X<br />

X<br />

i=1<br />

X


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 123<br />

For many C ∗ -<strong>algebras</strong> the tracial state space is determined completely by the<br />

K0-group, as we shall now explain. Let G be a pre-ordered abelian group and u an<br />

order unit in G, i.e. u ∈ G + is an element such that ∀g ∈ G ∃n ∈ N : g ≤ nu.<br />

A homomorphism ϕ : G → R is positive when ϕ(G + ) ⊆ [0, ∞[. A state on G is a<br />

positive element ϕ ∈ Hom(G, R) such that ϕ(u) = 1. The set of states of G will be<br />

denoted by SG. Note that the set of states of G depends not only on the partially<br />

ordered group structure of G, but also on the choice of order unit. In general there<br />

are many different order units in G and the corresponding state spaces may be very<br />

different. The state space SG of G comes equipped with a natural topology, the<br />

topology of pointwise convergence on G. Sets of the form<br />

{µ ∈ SG : |µ(gi) − ϕ(gi)| < ǫ, i = 1, 2, · · · , n}<br />

where ϕ ∈ SG, ǫ > 0 and the finite set {g1, g2, . . ., gn} ⊆ G may vary, form a basis<br />

for this topology.<br />

Lemma 4.49. SG is compact in the topology of pointwise convergence on G.<br />

Proof. Let [0, ∞] be the one-point compactification of [0, ∞[. By Tychonoffs<br />

theorem <br />

[0, ∞]<br />

g∈G<br />

is compact in the product topology. Recall that <br />

g∈G [0, ∞] consists of all functions<br />

f : G → [0, ∞]. We can therefore define Φ : SG → <br />

g∈G +[0, ∞] by Φ(ϕ)(g) =<br />

ϕ(g), g ∈ G + . We claim that<br />

Φ(SG) = {f ∈ <br />

[0, ∞] : f(g) + f(h) = f(g + h), g, h ∈ G + , f(u) = 1}. (4.9)<br />

g∈G +<br />

Obviously, Φ(SG) is contained in the set on the right-hand side. Conversely, if<br />

f ∈ <br />

g∈G +[0, ∞] such that f(g) + f(h) = f(g + h), g, h ∈ G + and f(u) = 1, then<br />

for any g ∈ G + , we can find n ∈ N such that g ≤ nu, i.e. such that g + h = nu for<br />

some h ∈ G + . Hence f(g) ≤ f(g) + f(h) = f(g + h) = f(nu) = nf(u) = n, proving<br />

that f(g) < ∞ for all g ∈ G + . It follows that we can define ω : G → R by<br />

ω(g − h) = f(g) − f(h), g, h ∈ G + .<br />

It is straightforward to check that ω ∈ SG and that Φ(ω) = f. This proves (4.9).<br />

Φ is then clearly an affine bijection between SG and {f ∈ <br />

g∈G +[0, ∞] : f(g) +<br />

f(h) = f(g + h), g, h ∈ G + , f(u) = 1}. It is straightforward to check that Φ is<br />

also a homeomorphism when the latter set is given the topology inherited from the<br />

product topology of <br />

g∈G +[0, ∞]. Since it is clearly closed in that topology the<br />

compactness of SG now follows from Tychonoff’s theorem. <br />

Note that SG is a convex subset of Hom(G, R) and that the topology just defined<br />

is the relative topology enherited from a locally convex vector space topology on<br />

Hom(G, R), a topology with a base consisting of sets of the form<br />

{µ ∈ Hom(G, R) : |µ(gi) − ϕ(gi)| < ǫ, i = 1, 2, · · · , n},<br />

where ǫ > 0, ϕ ∈ Hom(G, R) and the finite set {g1, g2, · · · , gn} ⊆ G may vary. Thus<br />

SG is a compact convex set.<br />

We now establish an important connection between the tracial state space T(A)<br />

of a unital C ∗ -algebra A and the state space SK0(A) of K0(A) (with respect to the


124 1. FUNDAMENTALS<br />

order unit [1]). Let ω ∈ T(A). We can then define, for each n ∈ N, a trace functional<br />

ωn on Mn(A) by<br />

<br />

ωn (aij) n<br />

= ω(aii) .<br />

In this way we can define a map ω∞ : P(A) → R by ω∞(p) = ωn(p) when p is a<br />

projection p in Mn(A). The trace property of each ωn ensures that ω∞(p) = ω∞(q)<br />

whenever p ≈ q. Furthermore, it is clear that ω∞(p ⊕ q) = ω∞(p) + ω∞(q), so we<br />

see that ω∞ gives rise to a semigroup homomorphism P(A)/≈ → R and hence, by<br />

the universal property of the Grotendieck construction, to a group homomorphism<br />

rA(ω) : K0(A) → R. It is straightforward to check that ωn is a positive functional<br />

on Mn(A), so we conclude that rA(ω)(K0(A) + ) ⊆ [0, ∞[, i.e. rA(ω) is positive. Furthermore,<br />

rA(ω)([1]) = 1, so we see that rA(ω) ∈ SK0(A). Thus we have produced<br />

a map<br />

rA : T(A) → SK0(A) .<br />

This map will be called the restriction map of A, because it is in some sense obtained<br />

by restricting ω to projections.<br />

i=1<br />

Lemma 4.50. The restriction map rA : T(A) → SK0(A) is continuous and affine.<br />

Proof. This is really straightforward, and we leave it as an exercise to the<br />

reader. <br />

Theorem 4.51. Let A be a unital C ∗ -algebra and A1 ⊆ A2 ⊆ A3 ⊆ · · · a<br />

sequence of unital C ∗ -sub<strong>algebras</strong> such that A = <br />

n An. Assume that the restriction<br />

map rAn : T(An) → SK0(An) is surjective for all n. It follows that rA : T(A) →<br />

SK0(A) is surjective.<br />

Proof. Let s ∈ SK0(A). We must construct ω ∈ T(A) such that rA(ω) = s.<br />

Let ιn : An → A be the inclusion map. Then s ◦ ιn∗ ∈ SK0(An), so the assumption<br />

on the rAn’s give us trace states ωn ∈ T(An) such that rAn(ωn) = s ◦ιn∗, n ∈ N. For<br />

each n we choose a state µn ∈ S(A) extending ωn : An → C. This is possible by the<br />

Hahn-Banach theorem. Since S(A) is compact there is a weak ∗ -accumulation point<br />

ω ∈ S(A) for the sequence {µn}. Let x, y ∈ A, ǫ > 0. Since <br />

n An is dense in A<br />

there is an N ∈ N and elements a, b ∈ AN such that xy −ab < ǫ and yx−ba < ǫ.<br />

Furthermore, since ω is an accumulation point for the µn’s, there is a k ≥ N such<br />

that |ω(xy) − µk(xy)| < ǫ and |ω(yx) − µk(yx)| < ǫ. Since µk is a trace state on AN<br />

we have that µk(ab) = µk(ba). Thus<br />

|ω(xy) − ω(yx)| ≤ |µk(xy) − µk(yx)| + 2ǫ ≤ |µk(ab) − µk(ba)| + 4ǫ = 4ǫ .<br />

Since ǫ > 0 was arbitrary this shows that ω(xy) = ω(yx), i.e. ω is a trace state.<br />

To show that rA(ω) = s it suffices to take a projection p ∈ Mk(A) for some k ∈ N<br />

and show that s([p]) = ω(p) , when ω is the extension of ω to a trace functional<br />

on Mk(A). To this end note that <br />

n Mk(An) is dense in Mk(A) so that there is an<br />

m ∈ N and a projection q ∈ Mk(Am) such that [p] = [q] in K0(A), cf. Lemma 3.7.<br />

For each l ≥ m we have that<br />

µl(q) = ωl(q) = rAl (ωl)([q]) = s([q]) .<br />

Since ω is a weak ∗ -accumulation point for the sequence {µn}, this implies that<br />

ω(q) = s([q]), i.e. that rA(ω)([p]) = s([p]).


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 125<br />

Corollary 4.52. Let A be a unital C ∗ -algebra and A1 ⊆ A2 ⊆ A3 ⊆ · · · a<br />

sequence of unital C ∗ -sub<strong>algebras</strong> such that A = <br />

n An. If each An has a trace<br />

state, then so does A.<br />

Proof. This follows from a straightforward modification of the proof of Theorem<br />

4.51 which we leave to the reader. <br />

Theorem 4.53. Let A be a unital C ∗ -algebra such that the span of projections<br />

in A is dense in A. Then the restriction map rA : T(A) → SK0(A) is injective.<br />

Proof. Let ω1, ω2 ∈ T(A) such that rA(ω1) = rA(ω2). For every projection<br />

p ∈ A we have that ω1(p) = rA(ω1)([p]) = rA(ω2)([p]) = ω2(p). Since A is the closed<br />

linear span of the projections in A it follows from the linearity and continuity of ω1<br />

and ω2 that ω1 = ω2. <br />

Corollary 4.54. Let A be a unital AF-algebra. Then the restriction map rA :<br />

T(A) → SK0(A) is an affine homeomorphism.<br />

Proof. By definition a unital AF-algebra is the closure of an increasing sequence<br />

of finite dimensional unital C ∗ -<strong>algebras</strong>. By Theorem 4.51, the surjectivity<br />

of rA : T(A) → SK0(A) follows if we can show that the restriction map of a finite<br />

dimensional C ∗ -algebra is surjective. This we leave to the reader. The injectivity<br />

of rA will follow from Theorem 4.53 if we can show that A is the closed linear span<br />

of its projections. But this is easy : If a = a ∗ ∈ A there is a b = b ∗ in some finite<br />

dimensional C ∗ -subalgebra of A such that a−b < ǫ. By linear algebra b is a linear<br />

combination of projections. <br />

Example 4.55. Let A be the complex vector space of sequences in M2 that<br />

converge to a diagonal element. More formally, let D2 denote the <strong>C∗</strong>-subalgebra of<br />

M2 consisting of the 2 × 2 matrices whose off diagonal elements are 0. Note that<br />

D2 ≃ C2 . Let<br />

A = {(xn) ⊆ M2 : lim xn ∈ D2} .<br />

n→∞<br />

Let Ak = {(xn) ∈ A : xn = xk+1, n ≥ k + 1}. Then Ak is a unital <strong>C∗</strong>-subalgebra of A isomorphic to C 2 ⊕ M2 ⊕ M2 ⊕ · · · ⊕ M2<br />

<br />

k times<br />

). Since A = <br />

k Ak, this shows that<br />

A is an AF-algebra. Note that K0(Ak) = Z k+2 as dimension groups and that the<br />

order unit [1] ∈ K0(Ak) corresponds to (1, 1, 2, 2, · · · , 2). Under this identification<br />

the inclusion map Ak → Ak+1 induces the map Z k+2 → Z k+3 given by<br />

(z1, z2, z3, · · · , zk+2) ↦→ (z1, z2, z3, · · · , zk+2, z1 + z2) .<br />

Thus K0(A) can be identified as the subgroup of Z N consisting of the elements<br />

(zn) ∈ Z N with the property that zk = z1 + z2 for all large enough k. We seek to<br />

identify SK0(A). To this end set<br />

pk = (1, 0, 0, 0, · · · , 0,<br />

1, 1, 1, · · ·)<br />

<br />

k times<br />

and<br />

qk = (0, 1, 0, 0, · · · , 0,<br />

1, 1, 1, · · ·) .<br />

<br />

k times<br />

Then pk, qk ∈ K0(A) and {pk} and {qk} are both decreasing sequences in K0(A).<br />

It follows that the limits s1 = limk→∞ s(pk) and s2 = limk→∞ s(qk) exist for any state


126 1. FUNDAMENTALS<br />

s ∈ SK0(A). Let rn, n = 3, 4, · · ·, be the element rn = (0, 0, 0, · · · 0, 0, 1, 0, 0, , · · · ) ∈<br />

1 at the n’th coordinate<br />

K0(A). Set sn = s(rn) for any state s ∈ SK0(A). We leave the reader to check that<br />

s (zn) ∞<br />

= snzn, (zn) ∈ K0(A)<br />

and<br />

n=1<br />

s1 + s2 + 2<br />

∞<br />

sn = 1<br />

n=3<br />

for all s ∈ SK0(A). Conversely, given any sequence {tn} in [0, 1] such that t1 + t2 +<br />

2 ∞ n=3 tn = 1, we can define a state s ∈ SK0(A) by<br />

s (zn) =<br />

∞<br />

i=1<br />

tizi .<br />

Then sn = tn for all n as the reader can easily verify.<br />

In this way we see that SK0(A) may be identified, as a convex set, with<br />

K = {(sn) ∈ [0, 1] N ∞<br />

: s1 + s2 + 2 sn = 1} .<br />

It is important to realize that this description of SK0(A), as a convex set, does not<br />

give any description of the topology of SK0(A). In particular, K is not a closed<br />

subseteq of [0, 1] N so SK0(A) is certainly not affinely homeomorphic to K when K<br />

is given the relative topology enherited from [0, 1] N . For example, 1<br />

2 rn ∈ K, n ≥ 3,<br />

and limn→∞ 1<br />

2 rn = (0, 0, 0, 0, · · ·) in the topology of [0, 1] N while limn→∞ 1<br />

2 rn =<br />

( 1 1<br />

, , 0, 0, · · ·) in the topology of K coming from the identification with SK0(A).<br />

2 2<br />

The topology of K which corresponds to the topology of SK0(A), and which makes<br />

the above identification of SK0(A) into an affine homeomorphism, is the topology of<br />

K given by pointwise convergence on K0(A) = {(zn) ∈ Z N : ∃N ∈ N such that zn =<br />

z1 + z2, n ≥ N}, i.e. (sα ∞ n) → (sn) in this topology if and only if limα i=1 sαi zi <br />

=<br />

∞<br />

i=1 sizi for all (zn) ∈ K0(A).<br />

By Corollary 4.54 the above description of SK0(A) is at the same time a description<br />

of T(A) as a compact convex set. Note that it is by no means apparent that<br />

T(A) is a Choquet simplex. Let us find Aff K and check that it is. To do this we first<br />

identify the extreme points of K = SK0(A). Define S1, S2 ∈ SK0(A) by Si (zn) <br />

=<br />

zi, i = 1, 2, and for k ≥ 3, define Sk ∈ SK0(A) by Sk (zn) = 1<br />

2zk, (zn) ∈ K0(A).<br />

We assert that<br />

∂eK ⊆ {Sk : k = 1, 2, 3, · · · }. (4.10)<br />

So let s be an extremal state on K0(A). Assume first that s(rk) = 0 for some k ≥ 3<br />

and define Pk : K0(A) → K0(A) by Pk (zn) = (0, 0, · · · 0, 0, zk, 0, 0, · · · ). Then<br />

zk at the k’th coordinate<br />

0 ≤ Pk(x) ≤ x for all x ∈ K0(A) + and hence 0 ≤ s ◦ Pk ≤ s. Set µ = s − s ◦ Pk.<br />

Then s = s ◦ Pk + µ , µ(u), s ◦ Pk(u) ∈ [0, 1] (where u = (1, 1, 2, 2, · · ·) is the order<br />

unit). Note that s ◦Pk(u) = s(2rk) > 0 and that µ(u)+s◦Pk(u) = 1. Now µ(u) = 0<br />

since otherwise<br />

s = µ(u)µ(u) −1 µ + s(2rk)s(2rk) −1 s ◦ Pk<br />

is a convex combination which yields that µ(u) −1 µ = s(2rk) −1 s ◦ Pk because s is<br />

extremal. This, however, is absurd since µ(rk) = 0 while s ◦ Pk(rk) = s(rk) > 0.<br />

n=3


4. CHOQUET SIMPLICES AND C ∗ -ALGEBRAS 127<br />

Thus µ(u) = 0 as asserted and hence µ = 0. It follows that s = s ◦ Pk and this<br />

clearly implies that s = Sk. Assume next that s(rk) = 0 for all k ≥ 3. In this case<br />

the above description of SK0(A) shows that there are numbers s1, s2 ∈ [0, 1] such<br />

that s (zn) = s1z1 + s2z2, (zn) ∈ K0(A). Note that s1 + s2 = 1. It is easy to see<br />

that s2 = 0 or s1 = 0 since s is extremal. In the first case s = S1 and in the second<br />

s = S2. We have established (4.10). The reader may check that the Sk’s are in fact<br />

extreme points in SK0(A) = K and hence that equality holds in (4.10). We shall<br />

not need this fact here.<br />

Note that we have that limk→∞ Sk = 1<br />

2S1 + 1<br />

2S2 in SK0(A). Indeed, for (zn) ∈<br />

<br />

K0(A) we have that Sk (zn) = 1<br />

2zk = 1<br />

2z1 + 1<br />

2z2 for all large enough k. Now<br />

define Φ : Aff SK0(A) → RN by Φ(f) = f(Sk) . Then φ maps Aff SK0(A) into<br />

{(tk) ∈ R N : limk→∞ tk = 1<br />

2 (t1 + t2) }. Denote this subspace of R N by X and<br />

set X+ = {(ti) ∈ X : ti ≥ 0 ∀i}. Then X is an ordered Banach space (in the<br />

norm (ti) = sup i |ti|) and 1 = (1, 1, 1, · · ·) is an order-unit. Note that Φ takes<br />

Aff SK0(A)+ into X+ and 1 to 1. If (ti) ∈ X+ we may define f ∈ Aff SK0(A)+ =<br />

Aff K+ by f (si) = t1s1 +t2s2 +2 ∞<br />

n=3 tnsn, (sn) ∈ K. Then Φ(f) = (ti), proving<br />

that Φ(Aff SK0(A)+) = X+. To prove that Φ is an isometry we shall need the<br />

following general fact<br />

When C is a compact convex set and f ∈ Aff C,<br />

then there is a an element c ∈ ∂eC such that |f(c)| = f.<br />

(4.11)<br />

Proof. Proof of (4.11) : Let x ∈ C such that |f(x)| = f. Set h = f if<br />

f(x) ≥ 0 and h = −f if f(x) < 0. Then h ∈ AffC and h(x) = h = f.<br />

F = {y ∈ C : h(y) = h} is then a compact convex non-empty subseteq of C. Let<br />

c be an extreme point of F. We assert that c ∈ ∂eC. To see this let a, b ∈ C, t ∈]0, 1[,<br />

such that c = ta+(1−t)b. Then h = h(c) = th(a)+(1−t)h(b) ≤ th+(1−t)h =<br />

h, proving that h(a) = h(b) = h, i.e. that a, b ∈ F. Since c is extreme in F we<br />

conclude that c = a = b, so that c extreme in C. Since |f(c)| = |h(c)| = h = f,<br />

we have proved (4.11). <br />

Given (4.11) and (4.10) it is obvious that Φ(f) = sup k |f(Sk)| = f, f ∈<br />

Aff SK0(A), i.e. Φ is is an isometry. Thus we have shown that Aff SK0(A) may<br />

be identified with X as an order unit space. We leave it as a (not necessarily easy)<br />

exercise for the reader to check that X has the Riesz interpolation property.<br />

The simplex K, with the topology coming from the identification with SK0(A), is<br />

qualitatively different from any example we have considered before. Observe namely<br />

that ∂eK is not closed in K since limk Sk = 1<br />

2S1 + 1<br />

2S2 in K. Therefore K can not<br />

be affinely homeomorphic to M1(X) for any compact Hausdorff space X.<br />

Example 4.56. Let A be a UHF-algebra. There is then a sequence A1 ⊆ A2 ⊆<br />

A3 ⊆ · · · of unital <strong>C∗</strong>-sub<strong>algebras</strong> of A such that each An is ∗- isomorphic to a<br />

matrix algebra and <br />

n An = A. By Lemma 4.45 each An has exactly one trace state<br />

ωn. Note that the uniqueness implies that<br />

ωn+1|An = ωn .<br />

We can therefore define ω : <br />

n An → C by ω|An = ωn, n ∈ An. Since |ω(x)| ≤ x<br />

for all x ∈ <br />

n An, we can extend ω by continuity to all of A. Then ω is a trace state<br />

on A. It follows easily from Lemma 4.45 that ω is the only trace state of A. Thus a


128 1. FUNDAMENTALS<br />

UHF-algebra has exactly one trace state. By Corollary 4.54 there must be one and<br />

only one state on K0(A). The reader is invited to check this.<br />

5.1. Ideals.<br />

5. Ideals and simple C ∗ -<strong>algebras</strong><br />

Definition 5.1. Let A be a C ∗ -algebra. An algebraic ideal in A is a linear<br />

subspace I ⊆ A such that a ∈ A, x ∈ I ⇒ ax, xa ∈ I. If I is also closed, I is called<br />

an ideal in A.<br />

Example 5.2. Let X be a compact Hausdorff space and A = C(X, Mk). Any<br />

closed subset F ⊆ X defines an ideal IF ⊆ A by<br />

IF = {f ∈ C(X, Mk) : f(x) = 0, x ∈ F } .<br />

The map F ↦→ IF defines a bijection between the closed subsets of X and the ideals<br />

on A = C(X, Mk). When I is an ideal in C(X, Mk) the corresponding closed subset<br />

of C(X, Mk) is given by<br />

{x ∈ X : f(x) = 0 ∀f ∈ I} .<br />

All unproven assertions of this example are left to the reader as exercises.<br />

For the study of ideals we need the notion of an approximate unit .<br />

Definition 5.3. Let I be an ideal in the C ∗ -algebra A. An approximate unit in<br />

I is a net {Eα} of positive elements in I such that<br />

1) Eα ≤ 1 for all α,<br />

2) α ≤ β ⇒ Eα ≤ Eβ,<br />

3) limα Eαx = x for all x ∈ I.<br />

Lemma 5.4. Let I be an ideal in the C ∗ - algebra A. Then I contains an approximate<br />

unit.<br />

Proof. Let U denote the set of finite collections (unordered tuples) α = {a1, a2, · · · , an}<br />

of elements from I. We introduce an ordering ≤ in U by setting α ≤ β when α =<br />

{a1, a2, · · · , an}, β = {b1, b2, · · · , bm} and there is an injective map σ : {1, 2, · · · , n} →<br />

{1, 2, · · · , m} such that ai = bσ(i) ∀i. Then U is a directed set. For α ∈ U as above<br />

we set<br />

n<br />

and<br />

Fα =<br />

i=1<br />

aia ∗ i<br />

Eα = nFα(1 + nFα) −1 ,<br />

where 1 ∈ Â. Then Eα ∈ I and Eα ≤ 1. To see that Eα ≤ Eβ when α ≤ β, note<br />

first that Eα = 1 − (1 + nFα) −1 . So to conclude that Eα ≤ Eβ we must show that<br />

(1 + mFβ) −1 ≤ (1 + nFα) −1 , where Fβ = m i=1 bib∗ i . Since α ≤ β we obviously have<br />

that nFα ≤ mFβ. But 1 + mFβ ≥ 1 + nFα ≥ 1 implies that X = (1 + nFα) −1 2(1 +<br />

mFβ)(1 + nFα) −1<br />

2 ≥ 1. Thus X−1 ≤ 1, i.e. (1 + nFα) 1<br />

2(1 + mFβ) −1 (1 + nFα) 1<br />

2 ≤ 1.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 129<br />

This implies that (1 + mFβ) −1 ≤ (1 + nFα) −1 , as needed. To see that limα Eαx = x<br />

for all x ∈ I, note that<br />

(Eαai − ai)(Eαai − ai) ∗ =<br />

i=1<br />

(1 + nFα) −1 Fα(1 + nFα) −1 =<br />

n<br />

(Eα − 1)aia ∗ i(Eα − 1) =<br />

F 1<br />

2<br />

α (1 + nFα) −2 F 1<br />

2<br />

α ≤ F 1<br />

2<br />

α (1 + nFα) −1 F 1<br />

2<br />

α =<br />

1<br />

n (1 − (1 + nFα) −1 ) ≤ 1<br />

1 .<br />

n<br />

Thus Eαai −ai2 ≤ 1<br />

n , i = 1, 2, · · · , n. It follows straightforwardly that limα Eαx =<br />

x for all x ∈ I.<br />

<br />

Theorem 5.5. An ideal I in a C ∗ -algebra A is selfadjoint (i.e. x ∈ I ⇒ x ∗ ∈ I)<br />

and A/I is a C ∗ -algebra in the quotient norm.<br />

Proof. Let {Eα} be an approximate unit in I. If x ∈ I, then limα x ∗ Eα−x ∗ =<br />

limα Eαx − x = 0, i.e. limα x ∗ Eα = x ∗ . Since x ∗ Eα ∈ I for all α, we see that<br />

x ∗ ∈ I, i.e. I is selfadjoint. In particular, A/I is a ∗-algebra and it suffices to prove<br />

the C ∗ -identity, i.e. that x + I 2 = xx ∗ + I. Since the reversed inequality is<br />

trivial, it suffices to show that x + I 2 ≤ xx ∗ + I. This depends on the fact that<br />

y + I = lim α y − Eαy, y ∈ A . (5.1)<br />

For all z ∈ I we have that y +z ≥ (1 −Eα)(y +z) for all α. Since limα Eαz = z,<br />

we see that y +z ≥ lim sup α y −Eαy. Hence y +I ≥ lim sup α y −Eαy. But<br />

Eαy ∈ I for all α, so clearly lim infα y − Eαy ≥ y + I, proving (5.1). Now we<br />

find that<br />

x + I 2 = lim α x − Eαx 2 = lim α (x − Eαx)(x − Eαx) ∗ =<br />

lim α (1 − Eα)xx ∗ (1 − Eα) ≤ lim α (1 − Eα)xx ∗ =<br />

lim α xx ∗ − Eαxx ∗ = xx ∗ + I .<br />

Example 5.6. Let A = C(X, Mk) be as in Example 5.2. If F ⊆ X is a closed<br />

subset and I = {f ∈ C(X, Mk) : f(x) = 0 ∀x ∈ F }, then A/I is ∗-isomorphic<br />

to C(F, Mk) and the quotient map can be identified with the restriction map r :<br />

C(X, Mk) → C(F, Mk) which restricts functions f ∈ C(X, Mk) to F. More formally,<br />

if q : A → A/I is the quotient map, there is a ∗-isomorphism α : A/I → C(F, Mk)<br />

such that<br />

commutes.<br />

Theorem 5.7. Let<br />

A1<br />

A r <br />

C(F, Mk)<br />

<br />

q <br />

<br />

<br />

α<br />

<br />

<br />

A/I<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·


130 1. FUNDAMENTALS<br />

be a sequence of C ∗ -<strong>algebras</strong>, A = lim<br />

−→ (An, ϕn) the corresponding inductive limit C ∗ -<br />

algebra. If I is an ideal in A,<br />

I = <br />

n<br />

ϕ∞,n(ϕ−1 ∞,n (I)) .<br />

Proof. Set In = ϕ−1 ∞,n (I) and let q : A → A/I be the quotient map. Then<br />

In = ker (q ◦ ϕ∞,n) and hence q ◦ ϕ∞,n induces an injective ∗-homomorphism qn :<br />

An/In → A/I in the obvious way, i.e defined such that qn(x + In) = q ◦ ϕ∞,n(x).<br />

Note that qn is injective and hence and isometry. Let x ∈ I and ǫ > 0. There is an<br />

n ∈ N and an element a ∈ An such that ϕ∞,n(a) − x < ǫ. Let b be the image of<br />

a in An/In and note that qn(b) = q ◦ ϕ∞,n(a) = q(ϕ∞,n(a) − x) < ǫ. Since<br />

qn is isometric this shows that inf{a − z : z ∈ In} < ǫ, i.e. there is an element<br />

a0 ∈ In such that a −a0 < ǫ. It follows that x −ϕ∞,n(a0) < 2ǫ. Since ǫ > 0 was<br />

arbitrary, we see that x ∈ <br />

n<br />

ϕ∞,n(ϕ−1 ∞,n(I)), proving that I ⊆ <br />

n<br />

The reversed inclusion is obvious. <br />

ϕ∞,n(ϕ −1<br />

∞,n(I)).<br />

The essence of this theorem is that an ideal of an inductive limit C ∗ -algebra is<br />

put together by ideals in the sequence of C ∗ -<strong>algebras</strong> defining the algebra. This is<br />

more transparent in the following corollary.<br />

Corollary 5.8. Let A be a C ∗ -algebra and A1 ⊆ A2 ⊆ A3 · · · an increasing<br />

sequence of C ∗ -sub<strong>algebras</strong> of A such that A = <br />

n An. If I is an ideal in A,<br />

I = <br />

An ∩ I .<br />

Proof. This follows straightforwardly from Theorem 5.7.<br />

n<br />

Lemma 5.9. Let F be a finite dimensional C ∗ -algebra and I an ideal in F. There<br />

is then a central projection q ∈ F such that I = qF.<br />

Proof. I is a finite dimensional C ∗ -algebra in itself, so I unital. Let q be the<br />

unit in I. It is then obvious that I = qF. To see that q is central in F, let x ∈ F.<br />

Since xq ∈ F and q is a unit in I, we have that qxq = xq. If x = x ∗ , this shows that<br />

qx = (xq) ∗ = (qxq) ∗ = qxq = xq. Hence q commutes with every selfadjoint element<br />

of F and therefore with every element of F, i.e. q is central.<br />

<br />

It is clear that if q is a central projection in any C ∗ - algebra A, then qA is an ideal<br />

in A. Lemma 5.9 shows that every ideal of a finite dimensional C ∗ -algebra is of this<br />

form. (The example A = C([0, 1]) shows that this is certainly not a general fact, cf.<br />

Example 5.6.) We know that a finite dimensional C ∗ -algebra F can be identified with<br />

a direct sum Mn1⊕Mn2⊕· · ·⊕Mnm of matrix <strong>algebras</strong>. The center of Mn1⊕Mn2⊕· · ·⊕<br />

) = ei, i = 1, 2, · · · , m.<br />

Mnm is spanned by the projections (0, · · · , 0, 0, 1, 0, 0, · · · , 0<br />

1 at the i’th entry<br />

In other words e1, e2, · · · , em are the minimal non-zero projections in the center of<br />

Mn1 ⊕ · · · ⊕ Mnm. The most general central projection is therefore of the form<br />

<br />

i∈D<br />

for some subset D ⊆ {1, 2, · · · , m}. So by using Lemma 5.9 we easily reach the<br />

following conclusion.<br />

ei


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 131<br />

Lemma 5.10. For any subset D ⊆ {1, 2, · · · , m}, define the ideal ID in Mn1 ⊕<br />

Mn2 ⊕ · · · ⊕ Mnm by<br />

ID = {(a1, a2, · · · , am) ∈ Mn1 ⊕ Mn2 ⊕ · · · ⊕ Mnm : ai = 0, i /∈ D} .<br />

This gives a bijection between the ideals of Mn1 ⊕ · · · ⊕ Mnm and the subsets on<br />

{1, 2, · · · , m}.<br />

Proof. Left to the reader. <br />

Lemma 5.11. Let I be an ideal in the C ∗ - algebra A. If y ∈ I, x ∈ A and<br />

0 ≤ xx ∗ ≤ y, then x ∈ I.<br />

Proof. After dividing through with y we may assume that y ≤ 1. By using<br />

this we prove that<br />

1<br />

lim y nx = x . (5.2)<br />

n→∞<br />

This will complete the proof because y 1<br />

n ∈ I for all n ∈ N. First we observe<br />

that t(1 − t 1<br />

n) converges decreasingly to 0 for all t ∈ [0, 1]. By Dinis theorem the<br />

convergence is uniform on [0, 1], proving that limn→∞ y(1 − y 1<br />

n) = 0. Since<br />

y 1<br />

nx − x 2 = (1 − y 1<br />

n)xx ∗ (1 − y 1<br />

n)<br />

≤ (1 − y 1<br />

n)y(1 − y 1<br />

n) ≤ y(1 − y 1<br />

n) ,<br />

(5.2) follows. <br />

Now we want to describe the ideals of an AF-algebra A. As one should expect<br />

the ideals can be read off from the K0(A)-group.<br />

Definition 5.12. Let G be a partially ordered group. An ideal H in G is a<br />

subgroup such that H = H ∩ G + − H ∩ G + and<br />

x ∈ G + , y ∈ H ∩ G + , x ≤ y ⇒ x ∈ H .<br />

Lemma 5.13. Let A be an AF-algebra and I an ideal in A. Set<br />

HI = {[e] − [f] : e, f is a projection in Mn(I) for some n ∈ N}<br />

Then HI is an ideal in K0(A).<br />

Proof. Note first that HI = HI ∩G + −HI ∩G + by definition. Let x ∈ G + , y ∈<br />

HI ∩ G + such that x ≤ y. Then y = [e] − [f] = [p] where e, f and p are projections<br />

, e, f in Mn(I) and p in Mn(A), for some n. By Lemma 3.27 1) we have that<br />

e ⊕ 0 ≈ p ⊕ f in M2n(A). Let v ∈ M2n(A) be a partial isometry implementing this<br />

equivalence. Since M2n(I) is an ideal in M2n(A) and e ⊕ 0 ∈ M2n(I), we see from<br />

Lemma 5.11 first that v ∈ M2n(I) and then that p, f ∈ Mn(I). Since x ∈ G + there<br />

is a projection q in some matrix algebra over A such that x = [q]. By increasing n<br />

we may assume that q ∈ Mn(A). Since [q] ≤ y = [p] there is a projection r ∈ Mn(A),<br />

orthogonal to q, such that q +r ≈ p in Mn(A), cf. Lemma 3.27 1). Since p ∈ Mn(I),<br />

it follows from Lemma 5.11 first that q +r ∈ Mn(I) and then that q ∈ Mn(I). Thus<br />

x = [q] ∈ HI. <br />

If we let i : I → A be the inclusion, the ideal HI of Lemma 5.13 is i∗ (K0 (I)).<br />

Let A be an AF-algebra and H is an ideal in K0(A). Set<br />

IH = {q ∈ A : q = q ∗ = q 2 , [q] ∈ H} .


132 1. FUNDAMENTALS<br />

We assert that span IH is an ideal in A. To see this, write A = <br />

n An where<br />

A1 ⊆ A2 ⊆ A3 ⊆ · · · are finite dimensional C ∗ -sub<strong>algebras</strong> of A. Let Jn = An ∩ IH.<br />

We assert that span Jn is an ideal in An. To see this, we may assume that<br />

An = Mn1 ⊕ · · · ⊕ Mnm,<br />

and let ek ij : i, j = 1, 2, . . ., nk, k = 1, 2, . . ., m be a full set of matrix units in An.<br />

Set<br />

D = k ∈ {1, 2, . . ., , m} : e k <br />

11 ∈ H .<br />

We prove first that<br />

span Jn = {(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D}. (5.3)<br />

To prove this equality, let q ∈ Jn = An ∩ IH. Write q = (q1, q2, . . ., qm). If qk = 0,<br />

we have that ek <br />

11 ≤ [gk] ∈ H which implies that ek <br />

11 ∈ H. It follows that k ∈ D.<br />

Thus q ∈ {(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D} and we conclude that span Jn ⊆<br />

{(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D}. To prove that the reverse inclusion, note that<br />

every element of {(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D} is a linear span of projections<br />

from {(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D}. It suffices therefore to show that any<br />

projection p ∈ {(a1, a2, . . .,am) ∈ An| ai = 0, i /∈ D} is in span Jn. To see that this<br />

is indeed the case, write p = (p1, p2, . . .,pm). If pk = 0, ek <br />

11 ∈ H and since<br />

k<br />

pk ≤ nk e11 , this implies that [pk] ∈ H. Hence pk ∈ IH, and we conclude that<br />

p ∈ span Jn.<br />

Note that (5.3) exhibits spanJn as an ideal in An, so to conlcude that span IH<br />

is an ideal in A it suffices now to show that<br />

span IH = <br />

span Jn. (5.4)<br />

Let q ∈ IH and let ǫ > 0. By Lemma ?? and Lemma 3.14 there is an n and<br />

a projection p ∈ An such that p and q are Murray-von Neumann equivalent and<br />

p − q ≤ ǫ. Then p ∈ An ∩ IH, so p ∈ <br />

n span Jn. Since ǫ > 0 was arbitrary, this<br />

implies that q ∈ <br />

n span Jn. It follows that span IH ⊆ <br />

n span Jn. Since the reverse<br />

inclsuion is trivial, we have established (5.4). Hence span IH is an ideal in A, as<br />

asserted.<br />

Theorem 5.14. Let A be an AF-algebra. The map I ↦→ HI is a bijection between<br />

the ideals of A and the ideals of K0(A). The inverse is the map H ↦→ span IH.<br />

Proof. We must show that span IHI = I for every ideal I in A and that<br />

Hspan = H for every ideal H in K0(A).<br />

IH<br />

I ⊆ span IHI : Let x ∈ I ∩ An. Then x is the span a set of projections in<br />

I ∩ An. Since the involved projections are in IHI , we see that x ∈ span IHI . Since<br />

I = <br />

n I ∩ An by Corollary 5.8<br />

Let first I be an ideal in A. If p ∈ IHI , [p] = [e] − [f], where e, f ∈ Mn(I)<br />

for some n. It follows that p ⊕ f ≈ e ⊕ 0 in Mn+1(A). Since e ⊕ 0 ∈ Mn+1(I), it<br />

follows from Lemma 5.11 that p ∈ I. Thus span IHI ⊆ I. To prove the converse<br />

inclusion, note that I is an AF-algebra in itself by Corollary 5.8 and Lemma 5.10.<br />

In particular I is the closed linear span of its projections, cf. the proof of Corollary<br />

4.54. It is therefore enough to prove that every projection e ∈ I is in IHI . But this<br />

is obvious. Hence span IHI = I.<br />

n


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 133<br />

Let next H be an ideal in K0(A) and take x ∈ Hspan ∩Σ(A). Then x = [e] for<br />

IH<br />

some projection e ∈ span IH. By using (??) we may in fact assume that e ∈ span Jn<br />

for some n ∈ N. But we saw above that every projection of span Jn is in IH, so<br />

x = [e] for some e ∈ IH, i.e. x ∈ H. Since Hspan IH is generated by Hspan ∩ Σ(A), IH<br />

we have that Hspan ⊆ H. If x ∈ H ∩ Σ(A), x = [e] for some projection e ∈ A.<br />

IH<br />

. <br />

Then e ∈ IH by definition and x ∈ H span IH . It follows that H ⊆ H span IH<br />

, with scale Σ(A), then H∩Σ(A) generates H∩K0(A) + as a semigroup. Indeed, if<br />

x ∈ H∩K0(A) + , there are elements z1, z2, · · · , zn ∈ Σ(A) such that z1+z2+· · ·+zn =<br />

x. In particular, 0 ≤ zi ≤ x, so zi ∈ H because H is an ideal in K0(A).<br />

5.2. Simple C ∗ -<strong>algebras</strong>. In this section we introduce one of the most interesting<br />

classes of C ∗ -<strong>algebras</strong>; namely the simple ones. We collect first a series of<br />

general facts about simple C ∗ -<strong>algebras</strong> and turn subsequently towards the study of<br />

simple AF-<strong>algebras</strong>.<br />

Definition 5.15. A C ∗ -algebra is simple when A has no non-trivial ideals, i.e.<br />

when {0} and A are the only ideals in A.<br />

Lemma 5.16. 1) A is simple if and only if the following condition holds :<br />

For every pair x, y ∈ A\{0} and every ǫ > 0 there are elements<br />

a1, a2, · · · , aN, b1, b2, · · · , bN ∈ A<br />

such that N<br />

i=1 aixbi − y < ǫ.<br />

2) When A is unital, A is simple if and only if A contains no non-trivial<br />

algebraic ideals.<br />

Proof. 1) : Assume first that the condition holds. If I is a non-zero ideal<br />

in A and x ∈ I\{0}, then the condition obviously implies that I is dense in A,<br />

i.e. is all of A. Thus A is simple. Conversely, assume that A is simple and let<br />

x, y ∈ A\{0}. The closure of span{axb : a, b ∈ A} is clearly an ideal which is<br />

non-zero because x ∗ xx ∗ x = 0 is in it. By simplicity of A it must be all of A, and<br />

hence y is in it. The existence of elements a1, a2, · · · , aN, b1, b2, · · · , bN in A such<br />

that <br />

i aixbi − y < ǫ is now almost immediate. 2) : If A does not contain<br />

any non-trivial algebraic ideals it is obviously simple. Conversely, assume that A is<br />

simple and let I be a non-zero algebraic ideal in A. Let x ∈ I\{0}. By 1) there<br />

are elements a1, a2. · · · , aN, b1, b2, · · · , bN ∈ A such that N<br />

i=1 aixbi − 1 < 1. But<br />

then N<br />

i=1 aixbi = z is an invertible element in I and hence 1 = zz −1 ∈ I. It follows<br />

that I = A. <br />

Example 5.17. A full matrix algebra Mn is simple. This follows from Lemma<br />

5.9 because 0 and 1 are the only central projections in Mn. (Check !) Now Corollary<br />

5.8 shows that any UHF-algebra is simple.<br />

Lemma 5.18. Let A be a simple C ∗ -algebra and ω : A → C a non-zero positive<br />

trace functional. Then ω is faithful , i.e. ω(a) > 0 for all a ∈ A+\{0}.<br />

Proof. Set I = {x ∈ A : ω(xx ∗ ) = 0}. Since |ω(ab)| 2 ≤ ω(aa ∗ )ω(bb ∗ ) for all<br />

a, b ∈ A, we see that I = {x ∈ A : ω(xy) = 0 ∀y ∈ A}. This shows that I is a<br />

closed subspace of A. Let x ∈ I, a ∈ A. Then ω(xay) = 0 and ω(axy) = ω(xya) = 0<br />

for all y ∈ A. Thus ax, xa ∈ I, i.e. I is an ideal in A. By assumption ω = 0, so


134 1. FUNDAMENTALS<br />

I = A, i.e. I = {0} by simplicity of A. In particular, ω(a) = ω(a 1<br />

2a 1<br />

2) > 0 for all<br />

a ∈ A+\{0}. <br />

Lemma 5.19. Let A be a simple C ∗ -algebra. Then Mn(A) is simple.<br />

Proof. Let {Eα} be an approximate unit in A. For each α we denote by E ij<br />

α<br />

the element of Mn(A) whose A-entries are all 0, except for the ij’th where the entry<br />

is Eα. Let I be a non-zero ideal in Mn(A) and x a positive non-zero element of I.<br />

If Eii αxEii α = 0 for all α and all i, it follows that Eii<br />

αx = 0 for all α and all i. Hence<br />

n<br />

x = limα i=1 Eii αx = 0, a contradiction. So, for some α and i, Eii αxE ii<br />

α is a non-zero<br />

positive element of I with only one non-zero entry, namely the ii’th. Since A is<br />

simple it follows easily from Lemma 5.16 1) that I contains all elements of Mn(A)<br />

whose only non-zero element is the ii’th. For each j = i and each b ∈ A, observe<br />

that<br />

diag(0, 0, · · · , 0, 0, b, 0, 0, · · · , 0, 0)<br />

=<br />

b at the j’th entry<br />

lim α E ji<br />

α<br />

diag(0, 0, · · ·0, 0, b, 0, · · · , 0, 0, 0)E<br />

b at the i’th entry<br />

ij<br />

α ∈ I .<br />

It follows that diag(a1, a2, · · · , an) ∈ I for all a1, a2, · · · , an ∈ A. In particular,<br />

Fα = diag(Eα, Eα, · · · , Eα) ∈ I for all α. Since {Fα} is an approximate unit in<br />

Mn(A), it follows that I = Mn(A). <br />

Lemma 5.20. Let A be a simple unital <strong>C∗</strong>- algebra and a ∈ A+ a non-zero<br />

positive element. There is then a finite set x1, x2, · · · , xN of element in A such that<br />

N<br />

= 1 .<br />

i=1<br />

xiax ∗ i<br />

Proof. By Lemma 5.16 there are elements c1, · · · , cn, d1, · · · , dn ∈ A such that<br />

n i=1 ciadi − 1 < 1. Then n i=1 ciadi is invertible. If we substitute each di with<br />

<br />

di j cjadj<br />

−1 n we get that i=1 ciadi = 1. Now, if x, y are arbitrary elements in A<br />

we have that xx∗ +y∗y ≥ xy+y ∗x∗ since (xx∗ +yy∗ )−(xy+y ∗x∗ ) = (x−y∗ )(x−y∗ ) ∗ ≥<br />

0. Therefore ciadi + d∗ iac∗i ≤ ciac∗ i + d∗i adi for each i and hence<br />

In particular,<br />

1 = 1<br />

2<br />

n <br />

ciadi +<br />

i=1<br />

n<br />

d ∗ iac ∗ 1<br />

i ≤<br />

2<br />

n <br />

i=1<br />

h = 1<br />

2<br />

n <br />

i=1<br />

ciac ∗ i +<br />

i=1<br />

n<br />

i=1<br />

ciac ∗ i +<br />

d ∗ <br />

iadi<br />

n<br />

i=1<br />

d ∗ <br />

iadi .<br />

is positive and invertible. Set xi = 1 √ h<br />

2 −1 2ci, i = 1, 2, · · · , n, and xn+i = 1<br />

√ h<br />

2 −1 2d∗ i, i =<br />

1, 2, · · · , n. Then 2n i=1 xiax∗ i = 1. <br />

Proposition 5.21. Let A be a unital simple C ∗ - algebra. Then any non-zero<br />

element of K0(A) + is an order-unit in K0(A).<br />

Proof. Since [1] is an order unit in K0(A) it suffices to show that if x ∈<br />

K0(A) + \{0}, then there are natural numbers n, m = 0 such that n[1] ≤ mx. Let<br />

x = [e] where e is a non-zero projection in Mn(A). By Lemma 5.19, B = Mn(A)


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 135<br />

is simple, so by Lemma 5.20 there are elements x1, x2, · · · , xm <br />

∈ B such that<br />

m<br />

i=1 xiex∗ i = 1B. Set<br />

⎛<br />

⎞<br />

x1e x2e . . . xme<br />

⎜ 0 0 . . . 0 ⎟<br />

v = ⎜<br />

⎝ .<br />

⎟<br />

. . .. .<br />

⎠<br />

0 0 . . . 0<br />

.<br />

Then v ∈ Mm(B) = Mmn(A) and vv ∗ = diag(1B, 0, 0, . . ., 0) ∈ Mm(B). Thus v is a<br />

partial isometry and v ∗ v must be a projection in Mm(B). Set E = diag(e, e, e, · · · , e)<br />

and note that vE = v. Hence v ∗ v = Ev ∗ vE ≤ E. It follows that n[1] =<br />

[diag(1B, 0, 0, · · · , 0)] = [v ∗ v] ≤ [E] in K0(A). Since [E] = m[e], this completes<br />

the proof. <br />

Proposition 5.22. Let A be a simple unital C ∗ - algebra. Then either K0(A) + =<br />

K0(A) or K0(A) + ∩ (−K0(A) + ) = {0}.<br />

Proof. Assume that K0(A) + ∩ (−K0(A) + ) = {0} and let 0 = x ∈ K0(A) + ∩<br />

(−K0(A) + ). By Proposition 5.21 x is an order unit in K0(A), so for any y ∈ K0(A)<br />

we can find m ∈ N such that y ≤ mx. Since mx ∈ −K0(A) + , this implies that<br />

y ≤ 0. Thus K0(A) = −K0(A) + , or equivalently, K0(A) = K0(A) + . <br />

Lemma 5.23. Let A be a simple C ∗ -algebra and q ∈ A a projection. Then qAq<br />

is a simple C ∗ -algebra.<br />

Proof. Let x, y ∈ qAq\{0}, ǫ > 0. Since A is simple there are elements<br />

a ′ 1 , a′ 2 , · · · , a′ N , b′ 1 , b′ 2 , · · · , b′ N<br />

∈ A such that <br />

i a′ i xb′ i − y < ǫ. Set ai = qa ′ i q, bi =<br />

qb ′ i q. Then ai, bi ∈ qAq for, all i and <br />

i aixbi − y < ǫ. <br />

We now turn to a study of the simple AF-<strong>algebras</strong>.<br />

Theorem 5.24. Let A be an AF-algebra. Then the following conditions are<br />

equivalent.<br />

1) A is simple.<br />

2) K0(A) has no ideals other than {0} and K0(A).<br />

3) Every non-zero element of K0(A) + is an order unit in K0(A).<br />

Proof. The equivalence between 1) and 2) is immediate from Theorem 5.14. 2)<br />

⇒ 3) : Let x ∈ K0(A) + \{0}. Set I = {y ∈ K0(A) : −mx ≤ y ≤ mx for some m ∈<br />

N}. It is easily seen that I is an ideal in K0(A), and it is non-zero since x ∈ I.<br />

Hence I = K0(A) by 2). But this means that x is an order unit in K0(A). 3) ⇒ 2)<br />

: Let I be a non-zero ideal in K0(A). Then I contains a non-zero element x from<br />

K0(A) + . By (3) x is a order unit, so for any y ∈ K0(A) there is an n ∈ N such that<br />

y ≤ nx and −y ≤ nx. Since I is an ideal in K0(A) and 0 ≤ nx − y ≤ 2nx ∈ I, we<br />

see that y = nx − (nx − y) ∈ I. Hence I = K0(A). <br />

For many purposes it is very useful to have a criterion for an AF-algebra to be<br />

simple which is given in terms of its Bratteli diagram. Therefore we introduce the<br />

following notation and terminology.<br />

Definition 5.25. Let A1 ⊕A2 ⊕ · · · ⊕An and B1 ⊕B2 ⊕ · · · ⊕Bm be direct sums<br />

of C ∗ -<strong>algebras</strong> and<br />

ϕ : A1 ⊕ A2 ⊕ · · · ⊕ An → B1 ⊕ B2 ⊕ · · · ⊕ Bm


136 1. FUNDAMENTALS<br />

a ∗-homomorphism. For each j ∈ {1, 2, · · · , n}, define ιj : Aj → A1 ⊕ A2 ⊕ · · · ⊕ An<br />

by<br />

ιj(a) = (0, 0, · · · , 0, 0, · · · , 0, 0, a, 0, 0, · · · , 0)<br />

.<br />

a at the j’th entry<br />

For each i ∈ {1, 2, · · · , m}, define πi : B1⊕B2⊕· · ·⊕Bm → Bi by πi(a1, a2, · · · , am) =<br />

ai. Then<br />

πi ◦ ϕ ◦ ιj : Aj → Bi<br />

is a ∗-homomorphism which we denote by ϕij. The ∗-homomorphisms ϕij, j =<br />

1, 2, · · · , n, i = 1, 2, · · · , m, will be called the partial maps of ϕ.<br />

Note that the partial maps determine ϕ since<br />

ϕ(a1, a2, · · · , an) = n <br />

ϕ1j(aj),<br />

j=1<br />

n<br />

ϕ2j(aj), · · · ,<br />

j=1<br />

n<br />

ϕmj(aj) <br />

Definition 5.26. A ∗-homomorphism ϕ : A1 ⊕ A2 ⊕ · · · ⊕ An → B1 ⊕ B2 ⊕<br />

· · · ⊕Bm is called mixing when ϕij = 0 for all i, j, i.e. when all its partials maps are<br />

non-zero.<br />

Lemma 5.27. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

be a sequence of finite direct sums of matrix <strong>algebras</strong>,<br />

An = Mk1 ⊕ Mk2 ⊕ · · · ⊕ Mkmn .<br />

Assume that for every k ∈ N there is an n > k such that ϕn,k : Ak → An is mixing.<br />

Then A = lim<br />

−→ (An, ϕn) is simple.<br />

Proof. Assume I is a non-zero ideal in A. By Theorem 5.7 ϕ −1<br />

∞,k (I) must be<br />

a non-zero ideal in Ak for some k ∈ N. Let qk ∈ Ak be a central projection such<br />

that qkAk = ϕ −1<br />

∞,k (I) and choose n1 > k such that ϕn1,k : Ak → An1 is mixing.<br />

Then ϕn1,k(qk) ∈ ϕ −1<br />

∞,n1 (I) and qϕn1,k(qk) = 0 for all non-zero central projections<br />

q in An1 because ϕn1,k is mixing. So Lemma 5.9 implies that ϕ−1 (I) = An1.<br />

∞,n1<br />

By assumption there is an n2 > n1 such that ϕn2,n1 : An1 → An2 is mixing. By<br />

repeating the argument we see that ϕ−1 ∞,n2 (I) = An2, and we get in this way a<br />

sequence n1 < n2 < n3 · · · in N such that Anm = ϕ−1 ∞,nm (I) for all m. Hence I = A<br />

by Theorem 5.7. <br />

In the general, non-unital case, the condition of Lemma 5.27 is only a sufficient,<br />

but not a necessary condition for a sequence of finite direct sums of matrix <strong>algebras</strong><br />

j=1


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 137<br />

to give a simple inductive limit C ∗ -algebra. For example, the Bratteli diagram<br />

1 1<br />

<br />

2 1<br />

<br />

3 1<br />

<br />

4 1<br />

. .<br />

describes such a sequence with limit K, the compact operators on l 2 , see Exercise<br />

2.27. The limit is simple but the condition of Lemma 5.27 is not satisfied. In the<br />

unital case, however, the situation is more satisfying.<br />

Theorem 5.28. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

be a sequence of finite direct sums of matrix <strong>algebras</strong>. Assume that the connecting<br />

∗-homomorphisms are unital and injective. Then A = lim(An,<br />

ϕn) is simple if and<br />

−→<br />

only if there is, for every n ∈ N, an m > n such that ϕm,n is mixing.<br />

Proof. Assume that A is simple and consider some n ∈ N. Let q1, q2, · · · , qmn<br />

be the minimal non-zero projections in the center of An. We must show that there<br />

is a m > n such that qϕm,n(qi) = 0 for all i = 1, 2, · · · , mn, and all non-zero central<br />

pojections q in Am. This will namely clearly imply that ϕm,n is mixing. Since A is<br />

simple we can find elements k j<br />

i , j = 1, 2, · · · , Ni, i = 1, 2, · · · , mn, in A such that<br />

Ni <br />

k<br />

j=1<br />

j j ∗<br />

iϕ∞,n(qi)k i = 1, i = 1, 2, · · · , mn .<br />

This follows from Lemma 5.20. But then there is an m ∈ N so large that Am contains<br />

such that<br />

elements r j<br />

i<br />

Ni <br />

<br />

j=1<br />

for all i = 1, 2, · · · , mn. In particular,<br />

r j j∗<br />

iϕm,n(qi)r i − 1 < 1<br />

Ni <br />

r<br />

j=1<br />

j j∗<br />

iϕm,n(qi)r i<br />

is invertible in Am and hence ϕm,n(qi)q must be non-zero for all non-zero central<br />

projections in Am.


138 1. FUNDAMENTALS<br />

5.3. Simple AF <strong>algebras</strong> and simple dimension groups.<br />

Definition 5.29. A unital C ∗ -algebra A is called approximately divisible when<br />

the following holds : For any finite subset F ⊆ A, any N ∈ N and any ǫ > 0, there is a<br />

finite-dimensional unital C ∗ - subalgebra B of A such that B ≃ Mn1⊕Mn2⊕· · ·⊕Mnm,<br />

where ni ≥ N, i = 1, 2, · · · , m, and a finite set G of A such that gb = bg for all<br />

g ∈ G, b ∈ B and<br />

∀f ∈ F ∃g ∈ G : f − g < ǫ .<br />

Less formally A is approximately divisible if any finite subset approximately<br />

commutes with a finite dimensional C ∗ -subalgebra which is the sum of matrix <strong>algebras</strong><br />

of arbitrary large dimensions. Maybe the definiton becomes a little more<br />

transparent if we introduce the following notation. When B is a C ∗ -subalgebra of<br />

the C ∗ -algebra A , set<br />

B ′ ∩ A = {a ∈ A : ab = ba ∀b ∈ B} .<br />

B ′ ∩A is a C ∗ -subalgebra of A and is called the relative commutant of B in A. Then<br />

A is approximately divisible when the following holds : For any finite subset F ⊆ A,<br />

any N ∈ N and any ǫ > 0, there is a finite dimensional unital C ∗ -subalgebra B ⊆ A<br />

such that<br />

where ni ≥ N, i = 1, 2, · · · , m, and<br />

B ≃ Mn1 ⊕ Mn2 ⊕ · · · ⊕ Mnm ,<br />

dist(x, B ′ ∩ A) < ǫ, x ∈ F.<br />

We are aiming to prove that all simple unital AF <strong>algebras</strong> which are not finite<br />

dimensional are approximately divisible. For this we need the following lemma.<br />

Lemma 5.30. Let A ≃ Mm and let B ≃ Mn be a unital C ∗ -subalgebra of A.<br />

Then<br />

B ′ ∩ A ≃ M m<br />

n .<br />

Proof. Let {eij} and {fij} be full sets of matrix units in A and B, respectively.<br />

By Lemma 2.6 there is a unitary u ∈ A such that<br />

ufiju ∗ =<br />

m<br />

n<br />

k=1<br />

ei+(k−1)n, j+(k−1)n<br />

for all i, j ∈ {1, 2, · · · , n}. The map x ↦→ uxu ∗ defines a ∗-isomorphism from B ′ ∩ A<br />

onto C ′ ∩ A where C = uBu ∗ . Note that {ufiju ∗ } form a full set of matrix units<br />

in C. By substituting C for B and ufiju ∗ for fij, we can therefore assume that<br />

fij =<br />

m<br />

n<br />

k=1 ei+(k−1)n,j+(k−1)n for all i, j. For each pair k, l ∈ {1, 2, · · · , m<br />

n<br />

pkl =<br />

n<br />

i=1<br />

ei+(k−1)n, i+(l−1)n .<br />

}, set<br />

It is straightforward, although possibly a little tedious, to check that {pkl} is a set<br />

of matrix units for a copy of M m<br />

n inside A and that fijpkl = pklfij for all choices<br />

of i, j, k, l. In other words, {pkl} are matrix units for a copy of M m<br />

n inside B′ ∩ A.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 139<br />

To see that {pkl} actually spans all of B ′ ∩ A, we first consider a, b ∈ {1, 2, · · · , m}.<br />

Straightforward calculations show that<br />

<br />

fijeabfji = <br />

ei+(k−1)n,j+(k−1)neabej+(l−1)n,i+(l−1)n = pkl ,<br />

i,j<br />

i,j,k,l<br />

where (k − 1)n + j = a and (l − 1)n + j = b, provided that a and b have the same<br />

remainder after division by n. If this is not the case we find that<br />

<br />

fijeabfji = <br />

ei+(k−1)n,j+(k−1)neabej+(l−1)n,i+(l−1)n = 0.<br />

i,j<br />

i,j,k,l<br />

In either case <br />

i,j fijeabfji ∈ span{pkl}. Now, if x ∈ B ′ ∩ A, then<br />

x = 1<br />

n<br />

<br />

i,j<br />

fijxfji<br />

since x commutes with each fij. Thus, by writing<br />

x = <br />

λa,beab,<br />

where λa,b ∈ C for all a, b ∈ {1, 2, · · · , n}, we find that<br />

x = 1 <br />

λa,bfijeabfji ∈ span{pkl} .<br />

n<br />

i,j,a,b<br />

a,b<br />

Hence B ′ ∩ A = span{pkl} ≃ M m.<br />

<br />

n<br />

In the following lemma we use the convention that M0 = 0. We remind the<br />

reader that a ∗-homomorphism ϕ : Mn1 ⊕ · · · ⊕ MnM → Mm1 ⊕ · · · ⊕ MmM is inner<br />

equivalent to a standard homomorphism given by an M × N matrix S = (sij) with<br />

entries from N, cf. Lemma 2.6. The matrix S is called the multiplicity matrix for<br />

ϕ. (Two homomorphisms ϕ, ψ : A → B between <strong>C∗</strong>-<strong>algebras</strong> A and B are called<br />

unitarily equivalent when there is a unitary u ∈ ˆ B such that Ad u ◦ ϕ = ψ.)<br />

Lemma 5.31. Let ϕ : A = Mn1 ⊕ · · · ⊕ MnN → B = Mm1 ⊕ · · · ⊕ MmM be a<br />

unital ∗-homomorphism. Then<br />

ϕ(A) ′ ∩ B ≃ ⊕<br />

i,j∈{1,2,···,M}×{1,2,···,N}<br />

where (sij) is the multiplicity matrix for ϕ.<br />

Proof. For each i ∈ {1, 2, · · · , M}, let πi : B → Mmi<br />

ϕ(A) ′ ∩ B ≃ ⊕ M i=1 πi ◦ ϕ(A) ′ ∩ Mmi .<br />

Msij ,<br />

be the projection. Then<br />

Fix i and let q1, q2, · · · , qN be the minimal non-zero central projections in A. Then<br />

pj = πi ◦ ϕ(qj), j = 1, 2, · · · , N, are orthogonal projections with sum 1 in Mmi . Set<br />

Dij = pjMmipj and Cij = πi ◦ ϕ ◦ ιj(Mnj ), where ιj : Mnj → A is as in Definition<br />

5.25. Then the map<br />

x ↦→ (xp1, xp2, · · · , xpN)<br />

defines a ∗-isomorphism from πi ◦ ϕ(A) ′ ∩ Mmi onto ⊕Nj=1 C′ ij ∩ Dij. Since Cij ≃<br />

Mnj and Dij ≃ Msijnj by definition of (sij), we conclude from Lemma 5.30 that<br />

C ′ ij ∩ Dij ≃ Msij . Hence<br />

ϕ(A) ′ ∩ B ≃ ⊕ M i=1πi ◦ ϕ(A) ′ ∩ Mmi ≃ ⊕Mi=1 ⊕Nj=1 C′ ij ∩ Dij ≃ ⊕ Msij<br />

ij<br />

.


140 1. FUNDAMENTALS<br />

Lemma 5.32. Let<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

be a sequence of finite dimensional <strong>C∗</strong>-<strong>algebras</strong> and unital injective ∗-homomorphisms.<br />

Assume that A = lim(An,<br />

ϕn) is simple and not finite dimensional. Let S<br />

−→ n,m = (s n,m<br />

ij )<br />

be the multiplicity matrix for ϕn,m : An → Am, n < m. It follows that<br />

for all n ∈ N.<br />

limm→∞ min<br />

ij sn,m ij<br />

= ∞<br />

Proof. Assume first that there is sequence k1 < k2 < · · · in N such that Akn is<br />

a matrix algebra for all n. Then A is the inductive limit of the sequence<br />

A1<br />

ϕ1 <br />

A2<br />

ϕ2 <br />

A3<br />

ϕ3 <br />

· · ·<br />

(and hence A is a UHF algebra). Since A is not finite dimensional we must have<br />

that limm→∞ dim Akm = ∞. Since Ski,kj is the one-by-one matrix with entry<br />

S ki,kj<br />

<br />

dim Akj 11 =<br />

, we have that limj→∞ S dim Aki ki,kj<br />

11 = ∞ for each i. Now, it is a general<br />

fact that S n,m = S k,m S n,k when n < k < m. Since the ϕj’s are unital and injective,<br />

so that no row or column is zero in any of the S n,m ’s, it follows that<br />

min<br />

ij Sn,m ij<br />

increases with m −n. Therefore we have that limm→∞ minij S n,m<br />

ij<br />

<br />

= ∞ in this case.<br />

In the remaining case there is an N ∈ N such that S k,k+1 has at least two rows<br />

when k ≥ N. By Theorem 5.28 there is a sequence k1 < k2 < · · · in N such that<br />

min<br />

ab Ski,kj<br />

ab ≥ 1<br />

for all i < j. Since S ki,ki+1 has at least two rows when ki ≥ N, it follows easily that<br />

min<br />

ab Ski,kj<br />

ab<br />

≥ 2j−i−1<br />

for all i < j with ki ≥ N. In particular, limj→∞ minab S ki,kj<br />

ab<br />

The conclusion that limm→∞ minij S n,m<br />

ij<br />

= ∞ when ki ≥ N.<br />

= ∞ follows now as in the first case. <br />

Theorem 5.33. Let A be a simple unital AF-algebra which is not finite dimensional.<br />

Then A is approximately divisible.<br />

Proof. Let F be a finite subset of A, N ∈ N and ǫ > 0. Write A = <br />

n An<br />

where A1 ⊆ A2 ⊆ A3 ⊆ · · · is an increasing sequence of finite dimensional unital<br />

<strong>C∗</strong>-sub<strong>algebras</strong> of A. There is then an n ∈ N and a finite set G ⊆ An such that<br />

dist(x, G) < ǫ for all x ∈ F. By Lemma 5.31 and Lemma 5.32 there is an m > n<br />

such that B = A ′ n ∩ Am ≃ Mn1 ⊕ · · · ⊕ Mnd , where ni ≥ N for all i = 1, 2, · · · , d.<br />

Since G ⊆ B ′ ∩ A, this completes the proof.<br />

<br />

Lemma 5.34. Let A be a simple unital AF-algebra which is not finite dimensional.<br />

Let p be a projection in A and α ∈ [0, 1]. For any ǫ > 0 there is a projection q ∈ A<br />

such that<br />

|αω(p) − ω(q)| < ǫ ∀ω ∈ T(A) .


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 141<br />

Proof. We may obviously assume that p = 0. Choose N ∈ N such that 1<br />

N<br />

Since A is approximately divisible by Theorem 5.33, there is a finite dimensional<br />

unital <strong>C∗</strong>-subalgebra B of A such that B ≃ Mn1 ⊕ · · · ⊕ MnL , where ni ≥ N for all<br />

i, and dist(p, B ′ ∩ A) is as small as we want. By using first Lemma 2.10 and then<br />

Lemma 2.11 we can produce a projection e ∈ B ′ ∩ A such that p ≈ e. In particular,<br />

ω(p) = ω(e) > 0, ∀ω ∈ T(A), where the last inequality uses Lemma 5.18 and that<br />

p = 0. Let {rd ij : i, j = 1, 2, · · · , nd, d = 1, 2, · · · , L} be a full set of matrix units in<br />

B and set ed = e nd<br />

j=1 rd jj , d = 1, 2, · · · , L. For each i we choose ji ∈ {1, · · · , ni}<br />

such that |α − ji<br />

ni<br />

| ≤ 1<br />

ni<br />

≤ 1<br />

N<br />

q = e<br />

< ǫ. Set<br />

j1<br />

k=1<br />

r 1 kk<br />

+ e<br />

j2<br />

k=1<br />

r 2 kk<br />

+ · · · + e<br />

jL<br />

k=1<br />

r L kk .<br />

Then q is a projection and we claim that it does the job. To check it, note first that<br />

via er k 1j for all j, k, so that<br />

for all j, k. It follows that<br />

er k 11 ≈ erk jj<br />

ω(er k jj ) = ω(erk 1<br />

11 ) = ω(ek), ω ∈ T(A),<br />

nk<br />

ω(<br />

for all d = 1, 2, · · · , L. Thus<br />

jd<br />

k=1<br />

er d jd<br />

kk ) = ω(ed), ω ∈ T(A),<br />

nd<br />

|αω(p) − ω(q)| = |αω(e) − ω(q)| = |<br />

|<br />

L<br />

(α − jd<br />

d=1<br />

nd<br />

) ω(ed)<br />

|ω(e) ≤<br />

ω(e)<br />

L<br />

αω(ed) −<br />

d=1<br />

L<br />

|α − jd<br />

d=1<br />

nd<br />

| ω(ed)<br />

ω(e)<br />

< ǫ<br />

L<br />

jd<br />

nd<br />

d=1<br />

L<br />

d=1<br />

ω(ed)| =<br />

ω(ed)<br />

ω(e)<br />

= ǫ .<br />

for all ω ∈ T(A). <br />

5.4. Alternative proof of Lemma 5.34. In this subsection we present a more<br />

elementary proof of Lemma 5.34 which, in particular, avoids the notion of approximate<br />

divisibility.<br />

Lemma 5.35. Let A = Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN be a finite direct sum of matrix<br />

<strong>algebras</strong>, and set m = mini ni.<br />

For every t ∈ [0, 1] there is a projection p ∈ A such that |t − τ(p)| ≤ 1 for every<br />

m<br />

trace state τ ∈ T(A).<br />

Proof. Let ωi be the trace state of Mni , i.e.<br />

ωi = 1<br />

Tr,<br />

ni<br />

where Tr is the usual trace obtained by adding the diagonal entries. Note that for<br />

each i, there is a projection pi ∈ Mni such that<br />

|t − ωi(pi)| ≤ 1<br />

≤ 1<br />

. (5.5)<br />

m<br />

ni<br />

< ǫ.


142 1. FUNDAMENTALS<br />

Then<br />

p = (p1, p2, . . .,pN)<br />

is a projection in A which we claim has the desired property. To check it, let<br />

τ ∈ T(A) be a trace state. Let qi = (0, 0, . . ., 0, 1, 0, . . ., 0) be the central projection<br />

of A corresponding to the unit in Mni . Set αi = τ(qi), and note that<br />

N<br />

αi = 1. (5.6)<br />

i=1<br />

We can define a positive trace functional on Mni<br />

such that<br />

Mni ∋ ai ↦→ τ (0, 0, . . ., 0, ai, 0, . . ., 0).<br />

Since the value of this trace on the unit of Mni is αi, we conclude that τ (0, 0, . . ., 0, ai, 0, . . ., 0) =<br />

αiωi(ai) for all ai ∈ Mni . Thus, when a = (a1, a2, . . ., aN) ∈ A, we find that<br />

N<br />

N<br />

τ(a) = τ (0, 0, . . .,0, ai, 0, . . ., 0) = αiωi(ai). (5.7)<br />

i=1<br />

i=1<br />

By using (5.5), (5.7), and (5.6) twice, this leads to the estimate<br />

<br />

<br />

N N <br />

<br />

<br />

|t − τ(p)| = αit − αiωi(pi) <br />

<br />

≤<br />

N<br />

N<br />

αi |t − ωi(pi)| ≤<br />

i=1<br />

i=1<br />

i=1<br />

i=1<br />

αi<br />

1<br />

m<br />

= 1<br />

m .<br />

Lemma 5.36. Let A be a finite dimensional <strong>C∗</strong>-algebra and B ⊆ A a unital <strong>C∗</strong>- subalgebra. Then B ′ ∩ A is a finite dimensional <strong>C∗</strong>-algebra, so B ′ ∩ A ≃ Mn1 ⊕<br />

Mn2 ⊕ · · · ⊕ MnN for some natural numbers ni. Set m = mini ni.<br />

For every projection p ∈ B and every t ∈ [0, 1] there is a projection q ∈ A such<br />

that<br />

|tτ(p) − τ(q)| ≤ 1<br />

for all trace states τ ∈ T(A).<br />

m<br />

(5.8)<br />

Proof. By Lemma 5.35 there is a projection e ∈ B ′ ∩ A such that<br />

|t − θ(e)| ≤ 1<br />

(5.9)<br />

m<br />

for all trace states θ ∈ T(B ′ ∩ A). Set q = pe, and note that q is a projection in A.<br />

To check that q has the desired property, let τ ∈ T(A). If τ(p) = 0, we find that<br />

0 ≤ τ(q) = τ(pep) ≤ τ(p) = 0,<br />

i.e. τ(q) = 0. Thus (5.8) holds in this case. Assume therefore that τ(p) > 0. Then<br />

is a trace state of B ′ ∩ A, 2 and hence<br />

B ′ ∩ A ∋ x ↦→<br />

<br />

<br />

<br />

τ (ep) <br />

t − <br />

τ(p) <br />

τ (xp)<br />

τ(p)<br />

≤ 1<br />

m .<br />

Since 0 < τ(p) ≤ 1, this implies (5.8). <br />

2 Check this; it is essential that p ∈ B.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 143<br />

Here is the advertized alternative proof of Lemma 5.34: Let A1 ⊆ A2 ⊆ A3 ⊆ . . .<br />

be a sequence of unital finite dimensional <strong>C∗</strong>-sub<strong>algebras</strong> of A whose union is dense<br />

in A. By Lemma 3.15 there is an i ∈ N and a projection e ∈ Ai such that e and p<br />

are Murray-von Neumann equivalent. It follows from Lemma 5.31 and Lemma 5.32<br />

that there is a j ≥ i such that A ′ i ∩ Aj ≃ Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN , where<br />

1<br />

mini ni<br />

< ǫ.<br />

It follows then from Lemma 5.36 that there is a projection q ∈ Aj such that<br />

|tτ(e) − τ(q)| < ǫ<br />

for all τ ∈ T (Aj). When ω ∈ T(A) is a trace state of A, its restriction to Aj is a<br />

trace state of Aj and hence we conclude that<br />

|tω(e) − ω(q)| < ǫ<br />

for every trace state ω ∈ T(A). Since ω(e) = ω(p) for every ω ∈ T(A), this proves<br />

Lemma 5.34.<br />

For any unital C ∗ -algebra A with a non-empty tracial state space T(A) we can<br />

define a map<br />

ρA : K0(A) → Aff T(A)<br />

by ρA(x)(ω) = rA(ω)(x), ω ∈ T(A), where rA : T(A) → SK0(A) is the restriction<br />

map introduced in Section 4.4. ρA is obviously a group homomorphism and,<br />

although it need not be injective, it does carry important information about the<br />

order structure of K0(A). In fact, as we shall see, it determines the order structure<br />

completely when A is a simple unital AF-algebra.<br />

Lemma 5.37. Let A be a simple unital AF-algebra which is not finite dimensional.<br />

Then ρA(K0(A)) is dense in Aff T(A).<br />

Proof. Set V = ρA(K0(A)). Then V + V ⊆ V and −V = V since K0(A) is a<br />

group and ρA a group homomorphism. By Lemma 5.34, αV ⊆ V when α ∈ [0, 1], so<br />

it follows that V is a closed vector subspace of Aff T(A). To prove that V = Aff T(A)<br />

it therefore suffices, by the Hahn-Banach theorem, to show that any l ∈ Aff T(A) ∗<br />

which satisfies that l(V ) = {0} must be zero. From Lemma 4.9 we know that<br />

l = t1s1 −t2s2 where t1, t2 ∈ R + and s1, s2 are states on Aff T(A). But then there are<br />

trace states ωi ∈ T(A) such that si(f) = f(ωi), f ∈ Aff T(A), i = 1, 2, by Theorem<br />

4.3. Since l vanishes on V and 1 = ρA([1]) ∈ V , we have that l(1) = t1 − t2 = 0, i.e.<br />

t1 = t2. If t1 = t2 = 0, we must have that s1 − s2 annihilates V . In particular,<br />

0 = s1(ρA(x)) − s2(ρA(x)) = ρA(x)(ω1) − ρA(x)(ω2) = rA(ω1)(x) − rA(ω2)(x)<br />

for all x ∈ K0(A). It follows that rA(ω1) = rA(ω2) and hence that ω1 = ω2 by<br />

Corollary 4.54. Thus s1 = s2, proving that l = 0. <br />

Lemma 5.38. Let A be a unital AF-algebra. If x ∈ K0(A) is not positive (i.e. if<br />

x /∈ K0(A) + ), then there is tracial state ω ∈ T(A) such that rA(ω)(x) ≤ 0.<br />

Proof. Write A = <br />

n An where A1 ⊆ A2 ⊆ A3 ⊆ · · · is a sequence of unital<br />

finite-dimensional C ∗ -<strong>algebras</strong>. Since x is not positive in K0(A) there is an N ∈ N<br />

and elements xn ∈ K0(An)\K0(An) + such that x = ιn∗(xn), n ≥ N. By identifying<br />

K0(An) with Z mn for some mn ∈ N and using that xn /∈ Z mn<br />

+ , it is easy to construct<br />

a state sn on K0(An) such that sn(xn) < 0. By Corollary 4.54 (or an elementary


144 1. FUNDAMENTALS<br />

consideration) there is a tracial state ωn ∈ T(An) such that rAn(ωn) = sn, n ≥ N.<br />

Let µn be a state extension of ωn to A and let ω be a weak ∗ - condensation point of the<br />

sequence {µn} in S(A). As in the proof of Theorem 4.51 it follows that ω ∈ T(A).<br />

We have then that rA(ω)(x) = limn→∞ rAn(ωn)(xn) = limn→∞ sn(xn) ≤ 0. <br />

Theorem 5.39. Let A be a simple unital AF-algebra. Then<br />

K0(A) + = {x ∈ K0(A) : ρA(x) > 0} ∪ {0} .<br />

Proof. That the last set is contained in the first follows immediately form<br />

Lemma 5.38 and does not require simplicity of A. To prove the reversed inclusion,<br />

let p be a non-zero projection in Mn(A) for some n ∈ N. Note that ρA([p])(ω) =<br />

ωn(p), ω ∈ T(A), where ωn : Mn(A) → C is given by ωn ((aij)) = n<br />

i=1 ω (aii). Note<br />

that ωn is a positive trace functional. By Lemma 5.19 Mn(A) is simple and hence<br />

ωn is faithful by Lemma 5.18. Hence ωn(p) > 0 since p = 0. <br />

Lemma 5.37 and Theorem 5.39 are not only results on simple AF-<strong>algebras</strong>, but<br />

also a result on the structure of simple dimension groups. To make this clear we<br />

reformulate it as follows.<br />

Theorem 5.40. Let G be a countable simple dimension group with order unit u.<br />

Assume that G is not cyclic (i.e. G = Z). Then the map ϕG : G → Aff SG given by<br />

has dense image in Aff SG and<br />

ϕG(g)(s) = s(g), s ∈ SG,<br />

G + = {g ∈ G : ϕG(g) > 0} ∪ {0} .<br />

Proof. By Theorem 3.45 there is a unital AF-algebra A such that G ≃ K0(A)<br />

as dimension groups with order units. By Theorem 5.24 A is simple and A can not<br />

be finite dimensional because then K0(A) ≃ Z. By Corollary 4.54 rA : T(A) →<br />

SK0(A) is an affine homeomorphism, so it follows from Lemma 5.37 and Theorem<br />

5.39 that ϕK0(A) : K0(A) → Aff SK0(A) has dense range and K0(A) + = {x ∈<br />

K0(A) : ϕK0(A)(x) > 0}∪{0}. The isomorphism G ≃ K0(A) induces an isomorphism<br />

Aff SG ≃ Aff SK0(A) of order unit spaces in the obvious way such that the diagram<br />

ϕG<br />

G ≃ K0(A)<br />

ϕ K0 (A)<br />

❄ ❄<br />

Aff SG ≃Aff SK0(A)<br />

commutes. The conclusion follows from this. <br />

Note that the map ϕG : G → Aff SG of Theorem 5.40 can be defined for any<br />

partially ordered abelian group G with order unit. Theorem 5.40 tells us that this<br />

map actually determines the order structure on G when G is a countable simple noncyclic<br />

dimension group. This is actually true in greater generality; G need not be a<br />

dimension group to reach this conclusion, it suffices that G is unperforated. At this<br />

point it is not clear that such a purely algebraic assertion has any signifigance for<br />

C ∗ -<strong>algebras</strong> because we have not yet related arbitrary partially ordered unperforated<br />

groups to C ∗ -<strong>algebras</strong>. Here we continue by showing that Theorem 5.40 has a natural<br />

converse, namely the following.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 145<br />

Theorem 5.41. Let G be a torsion free abelian group and ∆ a metrisable Choquet<br />

simplex. Let ϕ : G → Aff ∆ be a group homomorphism such that ϕ(G) is dense in<br />

Aff ∆ and contains the constant function 1. Set<br />

G + = {g ∈ G : ϕ(g) > 0} ∪ {0} .<br />

When ordered by G + , G becomes a simple non-cyclic dimension group such that<br />

SG ≃ ∆ for any choice of order unit u in ϕ −1 (1).<br />

Proof. Since ϕ(G) is dense in Aff ∆, there is a g ∈ G such that 1 −ϕ(g) < 1,<br />

i.e. G + = {0}. Let g ∈ G + \{0}. For any h ∈ G there is an n ∈ N such that<br />

ϕ(h) < nϕ(g), and hence h ≤ ng. This proves that g, and hence any non-zero<br />

element of G + , is an order unit for G. In particular, G = G + − G + (i.e. G is<br />

directed). Since it is clear from the definition that G + ∩ (−G + ) = {0}, we have<br />

shown that G is a simple partially ordered abelian group. If g ∈ G and ng ∈ G +<br />

for some n ∈ N, we have that ng = 0 or nϕ(g) > 0. If ng = 0 we must have that<br />

g = 0 since G is torsion free and if nϕ(g) > 0 we must have that ϕ(g) > 0. Thus<br />

g ∈ G + , proving that G is unperforated. To check the Riesz interpolation property,<br />

let g1, g2, h1, h2 ∈ G + such that gi ≤ hj for all i, j ∈ {1, 2}. Then ϕ(gi) 2ǫ, i = 1, 2, ω ∈ ∆ .<br />

In case 1) k+ = k + ǫ is an element of Aff ∆ such that k+(ω) − gi(ω) > ǫ and<br />

hi(ω) − k+(ω) > ǫ for i = 1, 2, ω ∈ ∆. In case 2) k− = k − ǫ is an element of Aff ∆<br />

such that k−(ω) − gi(ω) > ǫ and hi(ω) − k−(ω) > ǫ for i = 1, 2, ω ∈ ∆. By the<br />

density of ϕ(G) we can find d ∈ G such that k± − ϕ(d) < ǫ. Then d interpolates<br />

the g’s and the h’s.<br />

Let u ∈ ϕ −1 (1). Define Φ u : ∆ → SG by Φ u (ω)(g) = ϕ(g)(ω). It is clear that<br />

Φ u is an affine continuous map. It is injective because ϕ(G) is dense in Aff ∆. To<br />

prove that Φ u is also surjective, let s ∈ SG. If g ∈ ker ϕ, we have that<br />

−u ≤ ng ≤ u<br />

for all n ∈ N. Hence −1 ≤ ns(g) ≤ 1 for all n ∈ N, showing that s(g) = 0, i.e.<br />

ker ϕ ⊆ ker s. We can therefore define ˜s : ϕ(G) → R by ˜s ◦ ϕ = s. If f = ϕ(g),<br />

for some g ∈ G, and −m m < f < , then −mu ≤ ng ≤ mu in G, and hence<br />

n n<br />

|˜s(f)| = |s(g)| ≤ m.<br />

A simple continuity argument shows now that |˜s(f)| ≤ f for<br />

n<br />

all f ∈ ϕ(G). Since ˜s is a group homomorphism, the density of ϕ(G) in Aff ∆ shows<br />

that ˜s extends by continuity to a group homomorphism Aff ∆ → R which we also<br />

denote by ˜s. Being a continuous group homomorphism, ˜s must actually be linear.<br />

Since s is positive on G + and ϕ(G + ) is dense in Aff ∆+, we have that ˜s is positive.<br />

Since ˜s(1) = s(u) = 1, we have that ˜s is a state on Aff ∆. By Theorem 4.3 there is<br />

a ω ∈ ∆ such that ˜s(f) = f(ω), f ∈ Aff ∆. Then Φu (ω) = s.


146 1. FUNDAMENTALS<br />

Theorem 5.40 and Theorem 5.41 give a complete description of the (countable)<br />

non-cyclic simple dimension groups with order unit. Hence they provide us with<br />

a powerful tool for the study of simple unital AF-<strong>algebras</strong>. For example, we get<br />

immediately the following<br />

Corollary 5.42. Let ∆ be a metrizable Choquet simplex. There is then a simple<br />

unital AF-algebra A such that T(A) is affinely homeomorphic to ∆.<br />

Proof. Since ∆ is metrizable, Aff ∆ is separable (by Exercise 4.12). Let {xn}<br />

be a dense sequence in Aff ∆. We may assume that x1 = 1. Set<br />

G = spanQ{xn : n ∈ N}<br />

and let ϕ : G → Aff ∆ be the inclusion map. By Theorem 5.41 there is a simple<br />

dimension group with order unit (namely, G with order unit 1), such that SG ≃ ∆.<br />

By Theorem 3.45 there is an AF-algebra A such that K0(A) ≃ G. Since G is simple,<br />

A is simple by Theorem 5.24. Since T(A) ≃ SK0(A) by Corollary 4.54, the proof<br />

is complete. <br />

Remark 5.43. In a weak moment one might think that we could classify nonunital<br />

AF-<strong>algebras</strong> by determining the isomorphism class of the AF-<strong>algebras</strong> obtained<br />

by adding a unit, and in this way avoid all considerations regarding scales of<br />

non-unital AF-<strong>algebras</strong>. However, the implication A + ≃ B + ⇒ B ≃ A fails in<br />

general, also among AF-<strong>algebras</strong>. To give an example of this consider the following<br />

two subsets of the real line :<br />

X = { 1<br />

: n = 2, 3, 4, · · · } ∪ {0} ,<br />

n<br />

and<br />

Y = { 1<br />

: n = 1, 2, 3, · · · } .<br />

n<br />

Both spaces are locally compact in the relative topology inherited from R, in<br />

fact X is compact. The abelian <strong>C∗</strong>-<strong>algebras</strong> A = C(X) and B = C0(Y ) are both<br />

AF-<strong>algebras</strong>. The reader is invited to write down a Bratteli diagram for each. But<br />

they are certainly not isomorphic since one is unital and the other is not. However,<br />

: n = 1, 2, 3, · · · } ∪ {0}.<br />

A + and B + are both isomorphic to C(Z) where Z = { 1<br />

n<br />

Example 5.44. The purpose of this example is to give a procedure to produce<br />

examples of partially ordered abelian groups which are directed and unperforated,<br />

but do no have the Riesz interpolation property, and hence are not dimension groups.<br />

Let H be a any non-zero partially ordered torsionfree abelian group which is directed<br />

and unperforated. Set G = Z ⊕ H with positive semi-group G + = {(z, h) : z ><br />

0, h ∈ H + \{0}} ∪ {(0, 0)}. We leave the reader to check that G is directed and<br />

unperforated. To see that G does not have the Riesz interpolation property, let<br />

h1, h2 ∈ H + such that h1 < h2 (by which we mean that h1 ≤ h2 and h1 = h2). For<br />

example we could take any non-zero element h1 of H + and choose h2 = 2h1. Set<br />

x1 = (0, 0), x2 = (0, h1) and y1 = (1, h2), y2 = (2, h2). Then xi ≤ yj in G for all<br />

i, j. Assume z = (a, b) is some element in G such that xi ≤ z ≤ yj for all i, j. It is<br />

easy to check that z can not be any of the elements x1, x2, y1 or y2. Therefore, since<br />

z − x1 ∈ G + , we must have that a > 0. On the other hand, since y1 − z ∈ G + , we<br />

must also have that a < 1. Since a ∈ Z, this is impossible.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 147<br />

Exercise 5.45. Let A be a C ∗ -algebra and set<br />

I = {I : I is an ideal in A} .<br />

I is ordered by inclusion, i.e. I1 ≤ I2 ⇔ I1 ⊆ I2. Show that I is a lattice and that<br />

for I1, I2 ∈ I.<br />

I1 ∧ I2 = span I1I2 = I1I2 = I1 ∩ I2,<br />

I1 ∨ I2 = I1 + I2 = I1 + I2 ,<br />

Exercise 5.46. Let A be a C ∗ -algebra and x ∈ A an arbitrary element of A.<br />

Show that xAx ∗ is a C ∗ -subalgebra of A. Show next that<br />

A is simple ⇒ xAx ∗ is simple.<br />

5.5. On the automorphism group. In this section we shall see how the classification<br />

result (and the methods leading to it) can be used to get some insight into<br />

the structure of the automorphism group of AF-<strong>algebras</strong>. For any C ∗ -algebra A we<br />

let Aut A denote the set of ∗-automorphism of A, i.e. an element α ∈ Aut A is an<br />

injective and surjective ∗-homomorphism α : A → A. Aut A is a group under composition;<br />

when α, β ∈ Aut A, α ◦ β is also an automorphism. Furthermore, AutA is<br />

a topological group. The norm on the bounded operators on A, considered just as<br />

a Banach space, gives rise to a metric D on Aut A; viz.<br />

D(α, β) = sup{α(a) − β(a) : a ≤ 1 } .<br />

It is not difficult to show (just do it !) that the Aut A is complete with respect to<br />

this metric. The corresponding topology is called the uniform topology.<br />

However, there is another topology which is more important, namely the one<br />

with a basis consisting of all the sets<br />

Vβ,a,ǫ = {α ∈ Aut A : α(a) − β(a) < ǫ} ,<br />

where β varies over all elements of Aut A, a over all elements of A and ǫ over R + .<br />

It is straightforward to see that a net {αt} in AutA converges to α ∈ Aut A in this<br />

topology if and only if limt αt(a) = α(a) for all a ∈ A. It is easy to see that Aut A<br />

is a topological group in this topology, and unless something else is explicitly stated,<br />

we will consider Aut A equipped with this topology which we shall refer to as the<br />

topology of point-wise normconvergence. In case the <strong>C∗</strong>-algebra A is separable, which<br />

is the case we shall mostly be concerned with, the topology is actually metrizable,<br />

i.e. comes from a metric. Indeed, when A is separable we can select a sequence {an}<br />

which is dense in the unit ball of A, and define a metric d on Aut A by<br />

∞<br />

d(α, β) = 2 −n α(an) − β(an) + α −1 (an) − β −1 (an) .<br />

n=1<br />

We leave the reader to check that the topology defined by the metric d is indeed the<br />

topology of pointwise norm-convergence.<br />

Exercise 5.47. Prove that the metric d is complete.<br />

Example 5.48. Let A be the C ∗ -algebra of sequences, {an}, in C (or in another<br />

C ∗ -algebra if you want) such that limn→∞ an = = 0. Define, for each m ∈ N, an<br />

automorphism βm of A by<br />

βm({an}) = (am, a1, a2, a3, a4, · · · , am−1, am+1, am+2, · · · · · ·) .


148 1. FUNDAMENTALS<br />

Then limm→∞ βm(a) = λ(a) for all a = {an} ∈ A, where λ : A → A is the<br />

endomorphism given by<br />

λ({an}) = (0, a1, a2, a3, a4, · · · · · ·) .<br />

Note that λ is certainly not an automorphism (it not surjective!). This example<br />

shows that Aut A is not complete with respect to a metric d0 given by d0(α, β) =<br />

∞<br />

n=1 2−n α(xn) − β(xn) for some dense sequence {xn} in the unit-ball of A. This<br />

is why we introduce the additional terms involving the inverses in the definition of d<br />

above. Note however that d and d0 determine the same topology on Aut A, namely<br />

the topology of pointwise normconvergence.<br />

An automorphism α ∈ AutA is called inner when there is a unitary u ∈ A +<br />

such that α(a) = uau ∗ for all a ∈ A. Of course, when A is unital this happens if<br />

and only if we can find such a unitary u in A. For any unital algebra B, we let<br />

U(B) denote the group of unitaries in B, i.e. U(B) = {u ∈ B : uu ∗ = u ∗ u = 1}.<br />

There is a group homomorphism U(A + ) → Aut A which sends u ∈ U(A + ) to the<br />

automorphism given by a ↦→ uau ∗ and the inner automorphisms are those which lie<br />

in the image of this map. We denote the group of inner automorphisms by Inn(A).<br />

The closure of Inn(A) will be denoted by Inn(A), and an automorphism in Inn(A)<br />

will be called approximately inner.<br />

Lemma 5.49. Inn(A) ⊆ {α ∈ Aut A : α∗ = idK0(A) on K0(A)}.<br />

Proof. It is obvious that the right-hand side defines a subgroup of Aut A, and<br />

since ueu ∗ ≈ e whenever e is a projection and u a unitary, it is clear that it contains<br />

Inn(A). We assert that its a closed subgroup. So let {αt} be net in AutA converging<br />

to α ∈ Aut A, and assume that αt∗ = idK0(A) for all t. Let e, f ∈ Mn(A + ) be<br />

projections such that [e] − [f] ∈ K0(A). For each automorphism β ∈ Aut A there<br />

is a unique automorphism β + ∈ Aut A + extending β. In the usual way we may<br />

furthermore extend β + to an automorphism β + n of Mn(A + ). Onserve that<br />

lim t αt + n (a) = α+ n (a)<br />

for all a ∈ Mn(A + ). In particular there are is a t such that αt + n (e) − α+ n (e) <<br />

1<br />

3 , αt + n (f) − α+ 1<br />

n (f) < 3 . But then, by Lemma 3.14, α∗([e] − [f]) = [α + n (e)] −<br />

[α + n (f)] = [αt + n (e)] − [αt + n (f)] = αt∗([e] − [f]) = [e] − [f] in K00(A + ). It follows<br />

that {α ∈ Aut A : α∗ = idK0(A) on K0(A)} is a closed subgroup containing<br />

Inn(A). Consequently it also contains Inn(A). <br />

Set<br />

AutΣ(K0(A)) = {γ ∈ Aut K0(A) : γ(Σ(A)) = Σ(A) } .<br />

AutΣ(K0(A)) is a subgroup of Aut K0(A) and it is clear that the map Aut A ∋ α ↦→<br />

α∗ ∈ Aut K0(A) maps into AutΣ(K0(A)).<br />

Theorem 5.50. Let A be an AF-algebra. Then the following is a short exact<br />

sequence :<br />

0<br />

⊆ α→α∗ <br />

Inn(A) <br />

Aut A <br />

AutΣ(K0(A))<br />

Proof. Let first γ ∈ AutΣ(K0(A)). By Theorem 3.35 3) there is an automorphism<br />

α ∈ Aut A such that α∗ = γ. This shows that the map α ↦→ α∗ is<br />

surjective. Let next α ∈ Aut A be an automorphism such that α∗ = id on K0(A).<br />

<br />

0


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 149<br />

There is then, by Theorem 3.35 2) a sequence of unitaries, {un}, in A + such that<br />

α(a) = limn→∞ unau∗ n for all a ∈ A. But this means that α ∈ Inn(A). This proves<br />

that the kernel of the map α ↦→ α∗ is contained in Inn(A). By Lemma 5.49 we<br />

therefore have that this kernel is exactly Inn(A). <br />

We are going to give an alternative description of the subgroup Inn(A) of Aut A.<br />

To do this we need some lemmas.<br />

Lemma 5.51. Let F be a finite dimensional C ∗ -subalgebra of A + and set<br />

F ′ ∩ A = {x ∈ A : xf = fx , f ∈ F } .<br />

There is then a surjective positive linear map P : A → F ′ ∩ A such that P(x) =<br />

x, x ∈ F ′ ∩ A, and P ≤ 2.<br />

Proof. Let {e d ij : i, j = 1, 2, · · · , nd, d = 1, 2, · · · , N} be a full set of matrix<br />

units in F and set p = <br />

d,i ed ii which is a projection in A. Define P : A+ → A + by<br />

P(a) = (1 − p)a(1 − p) +<br />

N<br />

nd<br />

1 <br />

e<br />

nd<br />

d=1 i,j=1<br />

d ijaedji . (5.10)<br />

Since ed ∗ d<br />

ji = eij we see that a ≥ 0 ⇒ P(a) ≥ 0 so that P is a positive linear map.<br />

For each m, k, l we find that<br />

and<br />

N<br />

e m klP(a) = em nd<br />

1 <br />

kl e<br />

nd<br />

d=1 i,j=1<br />

d ijaedji P(A)e m kl =<br />

N<br />

nd<br />

1 <br />

e<br />

nd<br />

d=1 i,j=1<br />

d ijaedji emkl = 1<br />

= 1<br />

nm<br />

nm<br />

nm<br />

j=1<br />

nm<br />

j=1<br />

e m kj aem jl<br />

e m kj aem jl .<br />

Hence P(a) ∈ F ′ ∩ A + for all a ∈ A + . If x ∈ F ′ ∩ A + we find that<br />

P(x) = (1 − p)x(1 − p) +<br />

= (1 − p)x + <br />

d,i<br />

N<br />

nd<br />

1 <br />

e<br />

nd<br />

d=1 i,j=1<br />

d ijxedji e d iix = (1 − p)x + px = x .<br />

= (1 − p)x +<br />

N<br />

nd<br />

1 <br />

e<br />

nd<br />

d=1 i,j=1<br />

d ijedji x<br />

In particular, P maps A onto F ′ ∩A. Let a = a ∗ ∈ A with a ≤ 1. Since −1 ≤ a ≤ 1<br />

in A + , the positivity of P and the fact that P(1) = 1 implies that −1 ≤ P(a) ≤ 1.<br />

Hence P(a) ≤ 1. If a ∈ A is a general element with a ≤ 1 we write a = a1 +ia2<br />

where ai = a ∗ i and ai ≤ a ≤ 1, i = 1, 2. It follows that<br />

P(a) = P(a1) + iP(a2) ≤ P(a1) + P(a2) ≤ 2 ,<br />

proving that P ≤ 2. <br />

Lemma 5.52. Let A be an AF-algebra and F ⊆ A a finite dimensional C ∗ -<br />

subalgebra of A. There is then an increasing sequence A1 ⊆ A2 ⊆ · · · of finite<br />

dimensional C ∗ -sub<strong>algebras</strong> of A such that A1 = F and A = ∞<br />

n=1 An.


150 1. FUNDAMENTALS<br />

Proof. Let B1 ⊆ B2 ⊆ B3 ⊆ · · · be an increasing sequence of finite dimensional<br />

<strong>C∗</strong>-sub<strong>algebras</strong> of A such that A = ∞ n=1 Bn. Let ι : F → A and in : Bn → A denote<br />

the inclusion maps. By Lemma 3.34 1) there is an m ∈ N and a ∗-homomorphism<br />

ψ : F → Bm such that ι∗ = im ∗ ◦ ψ∗. By Lemma 3.34 2) there is a unitary u ∈ A +<br />

such that ι = Adu ◦ im ◦ ψ. It follows that F ⊆ uBmu∗ . Set A1 = F and An =<br />

uBm+n−2u∗ , n ≥ 2. It straightforward to check that the sequence A1 ⊆ A2 ⊆ · · ·<br />

has the stated properties. <br />

Lemma 5.53. Let A be an AF-algebra and F ⊆ A a finite dimensional C ∗ -<br />

subalgebra. It follows that<br />

is an AF-algebra.<br />

F ′ ∩ A = {x ∈ A : xf = fx , f ∈ F }<br />

Proof. We pick an increasing sequence A1 ⊆ A2 ⊆ · · · of finite dimensional<br />

<strong>C∗</strong>-sub<strong>algebras</strong> in A such that F = A1 and A = ∞ n=1 An, cf. Lemma 5.52. Note<br />

that the map P : A → F ′ ∩ A defined in Lemma 5.51 maps An onto F ′ ∩ An for all<br />

n. We assert that<br />

F ′ ∩ A = <br />

F ′ ∩ An . (5.11)<br />

n<br />

This will of course complete the proof. Let x ∈ F ′ ∩A and choose xn ∈ An such that<br />

limn→∞ xn = x. Then P(x) = limn→∞ P(xn) since P is continuous, cf. Lemma<br />

5.51. But P(xn) ∈ F ′ ∩ An for all n, so this proves (5.11). <br />

Lemma 5.54. Let u, v be unitaries in a C ∗ -algebra A such that u −v ≤ µ < 2.<br />

There is then a continuous path γ : [0, 1] → U(A) such that γ(0) = u, γ(1) = v and<br />

γ(t) − v ≤ µ, t ∈ [0, 1].<br />

Proof. Since uv ∗ − 1 = u − v ≤ µ < 2, the spectrum of uv ∗ is a subset of<br />

T ∩ {λ ∈ C : |λ − 1| ≤ µ} ⊆ T\{−1}. Let ϕ : T\{−1} →] − 1, 1[ be the continuous<br />

function such that ϕ(e πit ) = t, t ∈]−1, 1[. Then ϕ(uv ∗ ) = a is a self-adjoint element<br />

of A with spectrum in<br />

ϕ({λ ∈ T : |λ − 1| ≤ µ})<br />

such that e 2πia = uv ∗ . Define γ(t) = e πita v, t ∈ [0, 1]. Then γ is a continuous<br />

path of unitaries connecting u to v. Since the spectrum of e 2πita is contained in<br />

{λ ∈ T : |λ − 1| ≤ µ} for all t, we find that γ(t) − v = e 2πita − 1 ≤ µ for all<br />

t ∈ [0, 1]. <br />

Proposition 5.55. Let A be a unital AF-algebra. Then the unitary group U(A)<br />

is pathconnected.<br />

Proof. First observe that the unitary group of Mn is pathconnected. Indeed, if<br />

u is a unitary n × n-matrix there is a unitary w ∈ Mn such that wuw ∗ is a diagonal<br />

unitary, i.e. is of the form<br />

diag(λ1, λ2, · · · , λn)<br />

for some λi ∈ T. Since T is pathconnected there is a path of diagonal unitaries<br />

connecting this unitary to 1 ∈ Mn. Hence wuw ∗ can be connected to 1 via a<br />

continuous path γ of unitaries. But then w ∗ γw will be a continuous path of unitaries<br />

connecting u to 1. It follows straightforwardly that the unitary group of any finite<br />

dimensional C ∗ -algebra is pathconnected.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 151<br />

Let now 1 ∈ A1 ⊆ A2 ⊆ · · · be a sequence of finite dimensional <strong>C∗</strong>-sub<strong>algebras</strong> of A such that A = <br />

n An. Let u ∈ U(A). For any ǫ we can find n ∈ N and x ∈ An<br />

such that u − x ≤ ǫ. If just ǫ is small enough, x∗x − 1 = x∗x − u∗u < 1 so<br />

that w = x(x∗x) −1<br />

2 is defined. w is then a unitary in An and if ǫ is small enough<br />

we have that u − w < 2. By Lemma 5.54 there is a continuous path of unitaries<br />

in U(A) connecting u to w. Since the unitary group of An is pathconnected we also<br />

have a path of unitaries (in An) connecting w to 1. Putting the two paths together<br />

we get a continuous path of unitaries in A connecting u to 1. <br />

Theorem 5.56. Let A be an AF-algebra. Then Inn(A) is the pathconnected<br />

component of AutA containing idA, i.e. Inn(A) consists of the automorphisms α<br />

for which there is a continuous path γ : [0, 1] → Aut A such that γ(0) = α and<br />

γ(1) = idA.<br />

Proof. Assume first that that there is such a path γ connecting α ∈ Aut A to<br />

idA. Then, when e is a projection in Mn(A), we get a continuous path, pt = γ(t)(p),<br />

of projections in Mn(A) connecting α(p) to p. By uniform continuity there is a<br />

n ∈ N such that pt − ps < 1<br />

2<br />

when |t − s| < . By using Lemma 3.14 we conclude<br />

3 n<br />

that<br />

[α(p)] = [p 1<br />

n<br />

] = [p 2<br />

n<br />

] = [p 3<br />

n<br />

] = · · · = [p n−1]<br />

= [p]<br />

n<br />

in K0(A). It follows that α∗ = idK0(A) and hence α ∈ Inn(A) by Theorem 5.50.<br />

Conversely, let α ∈ Inn(A) and choose an increasing sequence A1 ⊆ A2 ⊆ · · · of<br />

finite dimensional <strong>C∗</strong>- <strong>algebras</strong> with dense union in A. Since α∗ = idK0(A) there is<br />

for each n a unitary un ∈ U(A + ) such that α|An = Ad un. This follows from Lemma<br />

3.34 2). We will construct, for each n, a continuous path of unitaries wn t , t ∈<br />

[ 1 1 , n+1 n ], in U(A+ ) such that w1 1 = 1, w11 = u<br />

2<br />

∗ 1 , wn1 = u<br />

n<br />

∗ n−1 , wn1 = u<br />

n+1<br />

∗ n , n ≥ 2, and<br />

Ad wn t ◦ α|An−1 = idAn−1 for all t ∈ [ 1 1 , ]. We do this by induction. Since the<br />

n+1 n<br />

argument which starts the induction is the same as the that used for the induction<br />

step, we only give the latter. So assume that we have constructed wn t , t ∈ [ 1 1 , n+1 n ].<br />

Since unau∗ n = un+1au∗ n+1 (= α(a)), for all a ∈ An, we see that u∗ n+1un ∈ A ′ n ∩ A + .<br />

By combining Lemma 5.53 and Proposition 5.55 we see that the unitary group of<br />

A ′ n ∩ A+ is pathconnected. So there is a continuous path, vt, t ∈ [ 1<br />

n+2 , 1 ], of n+1<br />

= vtu∗ n. It is<br />

unitaries in A ′ n ∩A + such that v 1<br />

n+1<br />

= 1 and v 1<br />

n+2<br />

= u ∗ n+1un. Set w n+1<br />

t<br />

straigthforward to check that wn+1 has the required properties. Now define a path<br />

γ : [0, 1] → Aut A of automorphisms by γ(t) = Ad wn 1 1<br />

t ◦α, t ∈ [ , ], n ∈ N, and<br />

n+1 n<br />

γ(0) = idA. Then γ is a continuous path of automorphisms connecting α to idA. <br />

Thus, for an AF-algebra A, the group Inn(A) is the pathcomponent of Aut A<br />

containing the identity automorphism. As the next lemma shows, this is actually<br />

the same as the connected component containing idA.<br />

Lemma 5.57. Let A be an AF-algebra. The connected component of AutA containing<br />

idA is the same as the pathconnected component of Aut A containing idA and<br />

equals Inn(A).<br />

Proof. For every projection p in A, set<br />

Autp(A) = {α ∈ AutA : α(p) ≈ p}.


152 1. FUNDAMENTALS<br />

It follows from Lemma 3.27 that {α ∈ Aut A : α∗ = idK0(A)} = <br />

p Autp(A).<br />

Note that Autp(A) is a subgroup of AutA : If α(p) ≈v p and β(p) ≈w p, then<br />

β −1 ◦ α(p) ≈ β −1 (v) β −1 (p) ≈ β −1 (w ∗ ) p. Let α ∈ Autp(A). By Lemma 3.14 the open<br />

set<br />

{β ∈ Aut A : α(p) − β(p) < 1<br />

3 }<br />

is contained in Autp(A) so we see that Autp(A) is an open subgroup of Autp(A).<br />

But then Autp(A) is automatically also closed; indeed we can choose representatives,<br />

say {gi}, for the right Autp(A)-cosets in AutA so that Aut A is the disjoint union<br />

<br />

i gi Autp(A). If g0 Autp(A) is the coset containing idA, we have that the comple-<br />

ment of Autp(A) is <br />

i=0 gi Autp(A), an open set. Thus Autp(A) is both closed and<br />

open. Therefore it contains the connected component, Aut 0 A, of Aut A containing<br />

idA. It follows that Aut 0 A ⊆ <br />

p Autp(A) = {α ∈ Aut A : α∗ = idK0(A)} =<br />

Inn(A). The pathconnected component of Aut A containing idA, call it Aut0 A, is<br />

of course contained in Aut 0 A so by using Theorem 5.50 and Theorem 5.56 we find<br />

that<br />

Inn(A) = Aut0 A ⊆ Aut 0 A ⊆ Inn(A) .<br />

Let us consider some examples. Let’s begin with a matrix algbra A = Mn.<br />

Lemma 5.58. Every automorphism of Mn is inner, i.e. an arbitrary automorphism<br />

α ∈ Aut Mn is of the form<br />

α(x) = uxu ∗ , x ∈ Mn .<br />

Proof. Let {eij : i, j = 1, 2, · · · , n} be the usual matrix units in Mn. Note that<br />

α(eii), i = 1, 2, · · · , n, are projections such that α(e11) ≈ α(e22) ≈ · · · ≈ α(enn)<br />

and <br />

i α(eii) = 1. It follows that Tr(eii) = 1 for all i. In particular, we find that<br />

n<br />

α(e11) ≈ e11. Let v ∈ Mn be a partial isometry such that vv∗ = α(e11), v∗v = e11.<br />

It is straightforward to check that u = <br />

i α(ei1)ve1i is a unitary in Mn such that<br />

ueklu ∗ = α(ekl) for all k, l, proving that α = Ad u. <br />

It follows from this lemma that the homomorphism Un ∋ u ↦→ Ad u is a surjection.<br />

It is easy to see that the kernel of this homomorphism is λ1, λ ∈ T, so<br />

that<br />

Aut Mn ≃ Un/T .<br />

On AutMn the uniform topology agrees with the topology of pointwise normconvergence,<br />

and the above isomorphism is a homeomorphism when Un/T is equipped<br />

the natural topology. Note that the AutΣ(K0(Mn)) consists of the automorphisms of<br />

Z which fixes n, i.e. AutΣ(K0(Mn)) is trivial. So for this rather trivial C ∗ -algebra the<br />

short exact sequence of Theorem 5.50 is degenerate; the quotient is trivial. Consider<br />

next a finite-dimensional C ∗ -algebra<br />

F = Mn1 ⊕ Mn2 ⊕ · · · ⊕ MnN .<br />

Let α ∈ Aut F. To see how α acts on F, let ZF denote the center of F; i.e.<br />

ZF = {x ∈ F : xy = yx , y ∈ F }. It is straightforward to see that ZF ≃ C N .<br />

Indeed, set<br />

pi = (0, 0, · · · , 0, 1, 0, · · · , 0)


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 153<br />

where the 1 occurs on the i’th entry. Then pi ∈ ZF, i = 1, 2, · · · , N, and every<br />

element in ZF is a linear combination of the pi’s. Note that each pi is a minimal<br />

non-zero projection in ZF. Since also α(pi) must be a minimal non-zero projection<br />

in ZF, we find that there is a permuation σ of {1, 2, 3, · · · , N} such that<br />

α(ei) = eσ(i) , i = 1, 2, · · · , N .<br />

Note that α(eiF) = eσ(i)F for each i. Since eiF = Mni and eσ(i)F = Mn σ(i) we see<br />

that<br />

nσ(i) = ni<br />

for all i = 1, 2, · · · , N. Thus σ is not an arbitrary permutation; indeed, when we<br />

partition {1, 2, · · ·n} = m<br />

j=1 Fj such that ni = nk ⇔ i, k ∈ Fj we have that<br />

σ(Fj) = Fj , j = 1, 2, · · · , m. (5.12)<br />

Conversely, when σ is a permutation of {1, 2, · · · , N} such that (5.12) holds, we can<br />

define an automorphism β ∈ Aut F by<br />

β(a1, a2, · · · , aN) = (aσ(1), aσ(2), · · · , aσ(N)) , (a1, a2, · · · , aN) ∈ F .<br />

Then β(ei) = eσ(i) for all i. In other words, what we have shown is that there is<br />

a homomorphism Aut F → {σ ∈ ΣN : σ(Fj) = Fj , j = 1, 2, · · · , m} and this<br />

homomorphism is not only surjective; it is split surjective.<br />

Lemma 5.59. α ∈ Aut F is in Inn(F) = Inn(F) if and only α acts trivially on<br />

ZF.<br />

Proof. An inner automorphism of F must obviously act trivially on ZF, and by<br />

continuity the same must be true for an automorphism in Inn(F). If α acts trivially<br />

on ZF we find that α(ei) = ei and hence α(eiF) = eiF for all i. But eiF = Mni<br />

and since every automorphism of Mni is inner by Lemma 5.58, there is a unitary<br />

for each i such that<br />

ui ∈ Mni<br />

α(a1, a2, · · · , aN) = (u1a1u ∗ 1 , u2a2u ∗ 2 , · · · , uNaNu ∗ N )<br />

for all (a1, a2, · · · , aN) ∈ F. But then u = (u1, u2, · · · , uN) is a unitary in F such<br />

that α(x) = uxu ∗ . <br />

So for a finite dimensional C ∗ -algebra F as above, the short exact sequence of<br />

Theorem 5.50 becomes<br />

0<br />

<br />

InnF<br />

<br />

Aut F<br />

where the quotient map<br />

<br />

{σ ∈ ΣN : σ(Fj) = Fj , j = 1, 2, · · · , m} <br />

0 ,<br />

Aut F → {σ ∈ ΣN : σ(Fj) = Fj , j = 1, 2, · · · , m}<br />

is split. Note that {σ ∈ ΣN : σ(Fj) = Fj , j = 1, 2, · · · , m} ≃ Σ#F1 × Σ#F2 ×<br />

· · · × Σ#Fm.<br />

Lemma 5.60. Let H be a dense subgroup of R and set H + = H ∩ R + . Let<br />

A be a unital AF-algebra such that (K0(A), K0(A) + ) ≃ (H, H + ). Then every<br />

automorphism of A is approximately inner.


154 1. FUNDAMENTALS<br />

Proof. According to Theorem 5.50 the assertion is equivalent the assertion that<br />

any scale preserving automorphism of K0(A) is trivial. By Lemma 3.36 an automorphism<br />

γ of K0(A) is scale preserving if and only if γ([1A]) = [1A]. Using the<br />

isomorphism (K0(A), K0(A) + ) ≃ (H, H + ) we can consider the γ as an automorphism<br />

of H instead. By Lemma 5.71 it must have the form γ(x) = tx for some<br />

t > 0. But since γ fixes a non-zero element; namely the image of [1A], we find that<br />

t = 1. <br />

Example 5.61. It should be noted that it crucial that we deal with a unital AFalgebra<br />

in Lemma 5.60. To see this, let H = Q, ordered in the usual way, and let<br />

Σ = Q + . By Theorem 3.47 there is an AF-algebra A with (K0(A), K0(A) + , Σ(A)) ≃<br />

(Q, Q + , Q + ). Every non-zero element q ∈ Q + defines a scale-preserving automorphism<br />

of Q by s ↦→ sq, and it is straightforward to use Lemma 5.71 to conclude that<br />

every scale-preserving automorphism of Q arises this way. So we see that<br />

AutΣ(K0(A)) ≃ {q ∈ Q : q > 0} .<br />

The group composition on the right-hand side is given by multiplcation. So A admits<br />

certainly a large class of automorphisms which are not approximately inner.<br />

Example 5.62. Let A be a UHF-algebra, i.e. A is the inductive limit of a<br />

sequence<br />

Mnk<br />

ϕk<br />

−−−→ Mnk+1 ,<br />

where nk|nk+1 and ϕk(a) = diag(a, a, · · · , a). Then K0(A) is isomorphic, as a partially<br />

ordered abelian group, to the inductive limit of the sequence<br />

Z z↦→ nk+1 z nk Define µk : Z → R by µk(z) = z<br />

nk<br />

−−−−−→ Z.<br />

and note that<br />

µk<br />

Z z↦→ nk+1nk z <br />

Z<br />

µk+1<br />

<br />

<br />

R<br />

commutes. It follows that the µk’s define a homomorphism from K0(A) onto<br />

HA = { z<br />

: z ∈ Z, k ∈ N} .<br />

nk<br />

This map is easily seen to be an isomorphism of partially ordered groups when HA is<br />

ordered by HA ∩ R + . Note that HA is dense in R if and only if limk→∞ nk = ∞, i.e.<br />

if and only if A is not finite dimensional. So we see from Lemma 5.58 and Lemma<br />

5.60 that every automorphism of a UHF-algebra is approximately inner.<br />

Example 5.63. Let Aα be the unital AF-algebra with (K0(Aα), K0(Aα) + , [1]) ≃<br />

(Z + αZ, (Z + αZ) ∩ R + , 1), cf. Section 0.2. Then every automorphism of Aα is<br />

approximately inner.<br />

Example 5.64. Let A be an abelian C ∗ -algebra. Let A + is also abelian and<br />

hence any inner automorphism of A is trivial. Thus also Inn(A) is trivial. So if in<br />

addition A is an AF-algebra, we see that the short exact sequence of Theorem 5.50<br />

is degenerate in a new way; the quotient map is an isomorphism for abelian AF<strong>algebras</strong>.<br />

This does not mean that Aut A is a small group for such an AF-algebra;


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 155<br />

only that automorphisms of A are determined by how they act on K0(A). In fact,<br />

Aut A can easily be enormous in this case. Consider for example the AF-algebra<br />

C(X) where X = {0} ∪ { 1<br />

n : n ∈ N}. For any n ∈ N, the symmetric group Σn acts<br />

by homeomorphisms of X via<br />

σ( 1<br />

) =<br />

k<br />

<br />

1<br />

k<br />

, k ≥ n + 1,<br />

,<br />

, k ∈ {1, 2, · · · , n}<br />

1<br />

σ(k)<br />

σ ∈ Σn. It follows that Aut C(X) contains a copy of Σn for any n, and hence a copy<br />

of any finite group. But in fact, more is true. Let G be a countable infinite group.<br />

Then C(X) is isomorphic to<br />

{f : G → C : ∀ ǫ > 0, ∃ F ⊆ G , λ ∈ C such that #F < ∞, |f(g)−λ| < ǫ , g /∈ F } .<br />

For each g ∈ G define βg ∈ Aut C(X) by<br />

βg(f)(h) = f(g −1 h) , h ∈ G .<br />

Then g ↦→ βg is an injective group homomorphism. So we see that Aut C(X) contains<br />

a copy of every countable group G.<br />

5.6. Examples.<br />

5.6.1. A few facts about continued fractions. Let a0, a1, a2, · · · be a sequence of<br />

real numbers such that ai = 0, i ≥ 1. For each n, set<br />

[a0, a1, a2, · · · , an] = a0 +<br />

a1 +<br />

a2 +<br />

i.e. [a0] = a0, [a0, a1] = a0+ 1<br />

a1 , [a0, a1, a2] = a0+ 1<br />

a1+ 1<br />

a 2<br />

1<br />

1<br />

1<br />

a3 + · · · + 1<br />

an<br />

,<br />

, [a0, a1, a2, a3] = a0+ 1<br />

a1+ 1<br />

a 2 + 1<br />

a 3<br />

and so on. Consider the case where an ∈ N for all n ∈ N, i.e. the an’s is a sequence<br />

of natural numbers with ai > 0, i ≥ 1.<br />

For each n write<br />

[a0, a1, · · · , an] = pn<br />

qn<br />

where pn, qn ∈ N are mutually prime.<br />

Then<br />

and<br />

Lemma 5.65. Write [a1, a2, · · · , an] = p<br />

q<br />

pn = a0p + q<br />

qn = p .<br />

where p, q ∈ N are mutually prime.<br />

Proof. We have that [a0, a1, · · · , an] = a0 + [a1, a2, · · · , an] −1 so that pn<br />

qn =<br />

a0+ q a0p+q<br />

= . The conclusion follows because a0p+q and p are mutually prime. <br />

p p<br />

and<br />

Lemma 5.66.<br />

for k ≥ 2.<br />

pk = akpk−1 + pk−2 ,<br />

qk = akqk−1 + qk−2 ,


156 1. FUNDAMENTALS<br />

Proof. We shall prove this by induction in k. It is straightforward to check that<br />

p0 = a0, q0 = 1, p1 = a0a1 + 1, q1 = a1, p2 = a0a1a2 + a0 + a2 and q2 = a1a2 + 1, so<br />

that the lemma is certainly true for k = 2. Now assume that the lemma holds for<br />

k < n, regardless of which sequence a0, a1, a2, · · · we are given. By Lemma 5.65 we<br />

have that pn = a0p + q, qn = p where p, q ∈ N are mutually prime and satisfies that<br />

p<br />

q = [a1, a2, · · · , an] .<br />

By using the induction hypothesis we have that<br />

p = anu1 + u2 , q = anv1 + v2<br />

where u1, v1 are mutually prime, u2, v2 are mutually prime and<br />

and<br />

u1<br />

v1<br />

u2<br />

v2<br />

So by using Lemma 5.65 we find that<br />

and<br />

= [a1, a2, · · · , an−1] ,<br />

= [a1, a2, · · · , an−2] .<br />

pn = a0(anu1 + u2) + anv1 + v2 = an(a0u1 + v1) + a0u2 + v2 = anpn−1 + pn−2<br />

qn = anu1 + u2 = anqn−1 + qn−2 .<br />

This completes the induction step and hence the proof. <br />

Note that the conclusion of Lemma 5.66 can be written as the matrix identity<br />

<br />

pn qn an 1 pn−1 qn−1<br />

=<br />

, (5.13)<br />

pn−1 qn−1 1 0 pn−2 qn−2<br />

<br />

p1 q1<br />

valid for n ≥ 2. Since det = 1 this shows that<br />

or<br />

It follows that<br />

pn<br />

qn<br />

p0 q0<br />

pnqn−1 − pn−1qn = (−1) n−1 , n ≥ 1 . (5.14)<br />

pn<br />

qn<br />

− pn−2<br />

qn−2<br />

− pn−1<br />

qn−1<br />

= (−1)n−1<br />

qnqn−1<br />

= (−1)n−1<br />

qnqn−1<br />

, n ≥ 1 . (5.15)<br />

+ (−1)n<br />

qn−1qn−2<br />

, n ≥ 1 .<br />

Since the qn’s are increasing (this follows from Lemma 5.66) we find that<br />

when n is even and<br />

pn<br />

qn<br />

pn<br />

when n is odd. Hence we get the relations<br />

for all n ≥ 1.<br />

p2n−2<br />

q2n−2<br />

qn<br />

< p2n<br />

q2n<br />

> pn−2<br />

qn−2<br />

< pn−2<br />

qn−2<br />

< p2n+1<br />

q2n+1<br />

< p2n−1<br />

q2n−1


Thus we see immediately that<br />

5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 157<br />

p2n<br />

lim<br />

n→∞ q2n<br />

p2n−1<br />

≤ lim<br />

n→∞ q2n−1<br />

. (5.16)<br />

However, it follows from the recursive formulas from Lemma 5.66 that<br />

k<br />

[<br />

qk ≥ 2 2 ] , k ≥ 1,<br />

to that limk→∞ qk = ∞. So from (5.15) we conclude that limn→∞ pn<br />

Therefore the two limits from (5.16) agree and we have that the limit<br />

qn<br />

− pn−1<br />

qn−1<br />

pn<br />

lim<br />

n→∞ qn<br />

exists. We denote this limit by [a0, a1, a2, a3, · · ·], or if the space permits, by<br />

a0 +<br />

a1 +<br />

a2 +<br />

Exercise 5.67. Prove that limn→∞ pn<br />

qn<br />

1<br />

1<br />

a3 +<br />

1<br />

1<br />

a4 + · · ·<br />

is irrational.<br />

.<br />

= 0.<br />

Theorem 5.68. Let t ∈ R + \Q. There is then a unique sequence a0, a1, a2, · · ·<br />

in N such that ai > 0, i ≥ 1, and<br />

t = [a0, a1, a2, a3, · · ·] = a0 +<br />

a1 +<br />

a2 +<br />

1<br />

1<br />

a3 +<br />

1<br />

1<br />

a4 + · · ·<br />

Proof. Set a0 = [t] and define rn ∈ R + recursively by r1 = (t − a0) −1 and<br />

rn+1 = (rn − [rn]) −1 , n ≥ 1. This is possible because t is irrational. Set an = [rn]<br />

for n ≥ 1. Then<br />

1<br />

t = a0 + 1<br />

= a0 +<br />

for all n. Note that<br />

r1<br />

when n is even and that<br />

a1 +<br />

= a0 +<br />

a2 +<br />

1<br />

a1 + 1<br />

r2<br />

1<br />

1<br />

a3 + · · · + 1<br />

= · · · · · ·<br />

rn<br />

= [a0, a1, · · · , an, rn+1]<br />

t = [a0, a1, · · · , an, rn+1] ≥ [a0, a1, · · · , an]<br />

t = [a0, a1, · · · , an, rn+1] ≤ [a0, a1, · · · , an]<br />

when n is odd. From the first inequality we find that<br />

t ≥ [a0, a1, a2 · · ·] ,<br />

.


158 1. FUNDAMENTALS<br />

and from the second we find that<br />

t ≤ [a0, a1, a2 · · ·] .<br />

This proves the existence part of the statement. To prove uniqueness, let a ′ 0, a ′ 1, a ′ 2, · · ·<br />

= [t] = a0.<br />

be another sequence in N\{0} such that t = [a ′ 0 , a′ 1 , a′ 2 , · · ·]. Then a′ 0<br />

Hence [a1, a2, a3, · · ·] = [a ′ 1 , a′ 2 , a′ 3 , · · ·] and consequently, for the same reason, a1 = a ′ 1<br />

must be the integer part of this common value. Consequently, [a2, a3, a4, · · ·] =<br />

[a ′ 2, a ′ 3, a ′ 4, · · ·] and thus also a2 = a ′ 2. And so on. We leave it to the reader to<br />

formalize an induction argument which gives that an = a ′ n for all n. <br />

Let α ∈ R\Q. We can then define a homomorphism ϕα : Z 2 → R by<br />

ϕα(z1, z2) = z1 + αz2 .<br />

Then ϕα is injective because α is irrational. We set G = Z 2 , G + = {(z1, z2) :<br />

z1 + αz2 ≥ 0}, Gα = ϕα(G) ⊆ R and G + α = {t ∈ ϕα(G) : t ≥ 0}. With these<br />

definitions ϕα : G → Gα becomes an isomorphism of partially ordered groups. Note<br />

that Gα is unperforated and totally ordered because R has these properties. Thus<br />

Gα (and hence also G) is a dimension group. Each Gα contains 1 as an order unit.<br />

Lemma 5.69. For each α ′ ∈ R\Q there is an element α ∈]0, 1[\Q<br />

such that<br />

2<br />

(Gα ′, G+<br />

α ′, 1) ≃ (Gα, G + α , 1)<br />

as partially ordered groups with order unit (i.e. there is a group isomorphism ψ :<br />

and ψ(1) = 1.<br />

Gα ′ → Gα such that ψ(G +<br />

α ′) = G + α<br />

Proof. Let β ∈ R\Q. Then (z1, z2) → (z1, −z2) defines an automorphism γ of<br />

G = Z 2 such that γ(1, 0) = (1, 0) and ϕ−β ◦γ = ϕβ. Hence ϕ−β ◦γ ◦ϕ −1<br />

β : Gβ → G−β<br />

is an isomorphism of partially ordered groups with order unit. In short Gβ ≃ G−β<br />

as partially ordered groups. Let now k ∈ Z. Define γ : Z 2 → Z 2 by γ(z1, z2) =<br />

(z1 + kz2, z2). Then γ is an automorphism of Z 2 such that ϕβ ◦ γ = ϕβ+k. So as<br />

above we conclude that Gβ ≃ Gβ+k. These two conclusions imply the lemma. <br />

Lemma 5.70. Let H be a subgroup of R which is not dense in R. It follows that<br />

H is cyclic, i.e. there is an element s ∈ H such that H = Zs.<br />

Proof. We may assume that H = 0 so that H contains some element t > 0.<br />

Since H is not dense there is an open interval ]a, b[, b > a, such that ]a, b[∩H = ∅.<br />

Set d = b −a and note that ]0, d[∩H = ∅. It follows that s = inf{t ∈ H : t > 0} ≥<br />

d > 0. Then s ∈ H. Indeed, if s /∈ H we can find, by definition of s, two elements<br />

t1, t2 ∈ H such that s < t2 < t1 < 2s. But then t1 − t2 ∈ H∩]0, s[, a contradiction.<br />

It follows that Zs ⊆ H. If t is some element in H which is not contained in Zs, we<br />

have that ks < t < (k + 1)s for some k ∈ Z and then (k + 1)s − t ∈ H∩]0, s[, a<br />

contradiction. Thus Zs = H. <br />

Lemma 5.71. Let H be a dense subgroup of R and β : H → R a homomorphism<br />

such that h ≥ 0 ⇒ β(h) ≥ 0. It follows that there is a t ≥ 0 such that β(g) =<br />

tg, g ∈ H.<br />

Proof. Define ˜ β : R → R by<br />

˜β(t) = sup{β(x) : x ∈ H, x < t} .


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 159<br />

Consider s, t ∈ R. When x, y ∈ H such that x < t, y < s, we have that x+y < t+s<br />

and hence β(x) +β(y) = β(x + y) ≤ ˜ β(s +t). By taking supremum over x and y we<br />

get that<br />

˜β(t) + ˜ β(s) ≤ ˜ β(s + t) . (5.17)<br />

Let instead z ∈ H, z < t+s. Since t−H is dense in R, there is an element x ∈ H<br />

such that 0 < t −x < (t+s) −z. Then y = z −x = z −t+t−x < z −t+t+s−z =<br />

s, y ∈ H and x + y = z. Consequently, β(z) = β(x) + β(y) ≤ ˜ β(t) + ˜ β(s), and<br />

by taking the supremum over z we get the inequality which is reversed to (5.17),<br />

and hence we conclude that ˜ β is a homomorphism. Note that x ≥ 0 ⇒ ˜ β(x) ≥ 0.<br />

Take an element h ∈ H such that h > 0. Then ˜ β(h) = th for some unique element<br />

t ≥ 0. For each n ∈ N we find that n˜ β( 1<br />

nh) = th, i.e. ˜ β( 1 t h) = h. It follows<br />

n n<br />

that ˜ β(qh) = qth for all q ∈ Q. Let x ∈ R be arbitrary. There are sequences<br />

{xn}, {yn} ⊆ Qh such that the xn’s increases, the yn’s decreases and both have x<br />

as limit. Then txn = ˜ β(xn) ≤ ˜ β(x) ≤ ˜ β(yn) = tyn for all n, so we conclude that<br />

˜β(x) = tx. <br />

Lemma 5.72. Let α, γ ∈]0, 1<br />

2 [\Q. Then Gα = Z + αZ and Gγ = Z + γZ are<br />

isomorphic as partially ordered abelian groups with order unit 1 if and only α = γ.<br />

Proof. Let β : Gα → Gγ be an isomorphism of partially ordered groups such<br />

that β(1) = 1. Since Gα ≃ Z 2 is not cyclic, it must be dense by Lemma 5.70.<br />

Then we can apply Lemma 5.71 to conclude that β(h) = th for some t > 0. Since<br />

β(1) = 1 we have that t = 1, proving that Z + αZ = Z + γZ. Define χ1 : Z → T<br />

and χ2 : Z → T by χ1(z) = e 2πizα and χ2(z) = e 2πizγ , respectively. Then χ1 and χ2<br />

are both injective (because α and γ are irrational) and by the conclusion we have<br />

just obtained their range in T is the same. Hence χ −1<br />

2 ◦χ1 defines an automorphism<br />

of Z and must therefore be either the identity of the map z ↦→ −z. It follows that<br />

e2πiα = e2πiγ or e2πiα = e−2πiγ , i.e. α ±γ ∈ Z. Since α, γ ∈]0, 1[,<br />

this is only possible<br />

2<br />

if α = γ. <br />

By Theorem 3.45 there is, for each α ∈]0, 1<br />

that (K0(Aα), K0(Aα) + , [1]) ≃ (Gα, G + α , 1). It follows from Lemma 5.72 that the<br />

unital AF-<strong>algebras</strong> Aα and Aβ only are isomorphic if α = β. In the following we<br />

will describe a Bratteli diagram for the AF-algebra Aα. Let<br />

α = [0, a1, a2, a3, · · ·]<br />

2 [\Q, a unital AF-algebra Aα such<br />

be the continued fraction expansion of α, cf. Theorem 5.68. For each n there is a<br />

positive homomorphism<br />

ϕn : Z 2 → Z 2<br />

given by the matrix <br />

an 1<br />

.<br />

1 0<br />

Note that we here consider Z 2 as an ordered group with the usual ordering, i.e.<br />

as a simplex group. For each n ≥ 1, write [0, a1, a2, a3, · · · , an] = pn<br />

qn where<br />

pn, qn ∈ N\{0} are mutually prime. Set un = (qn−1 , qn−2) ∈ Z 2 when n ≥ 3. It<br />

follows from Lemma 5.66 that ϕn is a positive order unit preserving homomorphism<br />

when n ≥ 3.<br />

ϕn : (Z 2 , un) → (Z 2 , un+1)


160 1. FUNDAMENTALS<br />

Lemma 5.73. There is an isomorphism of partially ordered groups with order<br />

unit,<br />

(Gα, 1) ≃ lim((Z<br />

−→ 2 , un), ϕn)) .<br />

Proof. Define θn : Z2 → Z2 to be the homomorphism given by<br />

It follows from (5.13) that<br />

−1 pn−1 qn−1<br />

pn−2 qn−2<br />

−1 pn−1 qn−1<br />

pn−2 qn−2<br />

= (−1) n<br />

.<br />

qn−2 −qn−1<br />

−pn−2 pn−1<br />

Define ψα : Z 2 → Gα by ψα(z1, z2) = z1α + z2. It follows from (5.13) that we have<br />

the following commuting diagram<br />

(Z 2 , un)<br />

ψn<br />

θn <br />

(Z 2 , (0, 1)) ψα<br />

<br />

(Z2 , un+1) θn+1 ψα<br />

<br />

2 (Z , (0, 1))<br />

<br />

(Gα, 1)<br />

<br />

(Gα, 1)<br />

of positive order-unit preserving maps for each n ≥ 3. Since the θn’s are injective<br />

we get clearly an injective homomorphism θ : lim<br />

−→ (Z 2 , ϕn) → Gα which takes the<br />

order unit coming from the un’s to 1 ∈ Gα. It suffices to check that θ maps the<br />

positive semigroup in lim(Z<br />

−→ 2 , ϕn) onto G + α , or alternatively that<br />

<br />

First note that<br />

n<br />

ψα ◦ θn(Z 2+ ) = G + α . (5.18)<br />

ψα ◦ θn(z1, z2) = (−1) n (α(qn−2z1 − qn−1z2) + (−pn−2z1 + pn−1z2)) .<br />

Since (−1) n (αqn−2 − pn−2) ≥ pn−2<br />

qn−2 qn−2 − pn−2 = 0 and (−1) n (pn−1 − αqn−1) ≥<br />

pn−1 − pn−1 = 0 when n is even, we see that we have the inclusion ⊆ in (5.18).<br />

qn−1<br />

Conversely, let (z1, z2) ∈ Z2 such that αz1 + z2 > 0. Since limn→∞ pn<br />

= α, there is<br />

qn<br />

an N so large that pn−1z1 + qn−1z2 = qn−1( pn−1<br />

qn−1 z1 + z2) > 0 for all n > N. Hence<br />

θn −1 <br />

pn−1 qn−1<br />

(z1, z2) =<br />

(z1, z2) ∈ Z<br />

pn−2 qn−2<br />

2+<br />

for all n > N +1. Since αz1 +z2 = ψα ◦θn ◦θn −1 (z1, z2), we conclude that αz1 +z2 ∈<br />

<br />

n ψα ◦ θn(Z 2+ ) . <br />

So we see that Aα ≃ lim<br />

−→ (Fn, λn) where<br />

Fn = Mqn−2 ⊕ Mqn−1 , n ≥ 3<br />

and the connecting homomorphisms λn : Fn → Fn+1 is given by the Bratteli diagram<br />

qn−2 qn−1<br />

<br />

<br />

<br />

<br />

<br />

...<br />

qn−1 qn<br />

where there are an arrows from qn−1 to qn.<br />

<br />

.


5. IDEALS AND SIMPLE C ∗ -ALGEBRAS 161<br />

Exercise 5.74. Let A be a unital C ∗ -algebra which is approximately divisible.<br />

Show that ρA(K0(A)) is a subspace of Aff T(A). Show by example that this is not<br />

true in general.<br />

Exercise 5.75. Let A be a <strong>C∗</strong>-algebra and a ∈ A an element of norm a ≤ 1.<br />

Show that<br />

lim<br />

n→∞ (aa∗ ) 1<br />

n a = a.<br />

Exercise 5.76. Let H be an infinite-dimensional separable Hilbert space, and<br />

denote by L(H) the C ∗ -algebra of bounded operators on H.<br />

An operator A ∈ L(H) has finite rank when the range space, A(H), of A is finite<br />

dimensonal.<br />

a) Let A ∈ L(H) have finite rank. Show that A ∗ has finite rank.<br />

b) Let F be the set of operators in L(H) of finite rank. Show that F is an<br />

algebraic ∗-ideal.<br />

We denote by K the closure K = F of F in L(H). An element of K is a compact<br />

operator. By solving, or at least reading, the next problem, the reader will understand<br />

the reason for this name. It is not necessary to solve c) in order to do d)-g) -<br />

it is only included to explain the name.<br />

c) Let B be the unit ball in H, i.e. B = {ψ ∈ H : ψ ≤ 1}. Show that<br />

A ∈ L(H) is compact if and only if A(B) is a compact subset of H.<br />

d) Show that K is an ideal in L(H) which is not all of L(H).<br />

e) Let ei, i = 1, 2, 3, . . ., be an orthonormal basis in H, and define πn : Mn (C) →<br />

L(H) such that<br />

<br />

0, k ≥ n + 1,<br />

πn ((aij))ek = n i=1 aikei, k ≤ n.<br />

Show that πn is an injective ∗-homomorphism, and that πn (Mn (C)) ⊆ F.<br />

f) Show that K is an AF-algebra, and find a Bratteli diagram presenting it.<br />

g) Determine K0(K) as a scaled ordered group.<br />

Exercise 5.77. Let A be the unital AF-algebra given by the following Bratteli<br />

diagram:<br />

•<br />

<br />

<br />

<br />

<br />

• • •<br />

<br />

<br />

•<br />

<br />

<br />

<br />

• • •<br />

<br />

<br />

•<br />

<br />

<br />

<br />

• • •<br />

<br />

<br />

•<br />

<br />

<br />

<br />

• • •<br />

<br />

<br />

•<br />

<br />

<br />

<br />

. . . .


162 1. FUNDAMENTALS<br />

a) Show that A has three non-trivial ideals, and describe the inclusion pattern<br />

among the ideals in A.<br />

b) Let I1 and I2 be the maximal non-trivial ideals of A. Show that A/I1 and<br />

A/I2 are isomorphic AF-<strong>algebras</strong>.<br />

c) Let I0 be the minimal non-trivial ideal in A. Show that I0 is isomorphic to a<br />

well-known simple AF-algebra.<br />

d) Describe the simplex T(A) of trace-states of A.


CHAPTER 2<br />

Simple AI-<strong>algebras</strong><br />

163


CHAPTER 3<br />

The range of the Elliott invariant for simple unital<br />

AI-<strong>algebras</strong><br />

165

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!