• Keine Ergebnisse gefunden

REPRESENTATION THEORY OF C*-ALGEBRAS

N/A
N/A
Protected

Academic year: 2022

Aktie "REPRESENTATION THEORY OF C*-ALGEBRAS"

Copied!
27
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

REPRESENTATION THEORY OF C*-ALGEBRAS

Contents

1. Introduction 2

2. Basics 3

2.1. Operators on Hilbert spaces 3

2.2. Banach algebras 4

2.3. C*-Algebras 5

2.4. Unital Commutative C*-algebras 6

3. Further Structure 8

3.1. Unitization 8

3.2. Commutative C*-algebras 10

3.3. positive Elements 11

3.4. Approximate Units 13

4. Representation Theory of C*-Algebras 16

4.1. Representations 16

4.2. Positive Functionals, States 18

4.3. The GNS Construction 20

4.4. The Space of States, Pure States 22

4.5. The Theorem of Gelfand-Naimark 24

References 27

1

(2)

1. Introduction

When talking about Banach algebras one often has in mind the common ex- ample of bounded operators on a Hilbert space. However these operator algebras have more structure than just that of a Banach algebra - for example the adjoint operation. C*-algebras are Banach algebras that in addition have such an adjoint operation. The major part of this work will be concerned with showing that the Banach algebra structure together with this adjoint operation characterizes opera- tor algebras. In other words we will show that every C*-algebra is isomorphic to an algebra of operators (that is closed under adjoints) on a Hilbert space.

Section 2 reviews basic results and denitions concerning operators on Hilbert spaces, Banach algebras and C*-algebras. Most important the representation the- ory for commutative unital C*-algebras will be discussed. The results of Section 2 will be given without proofs. For proofs and further results the reader is referred to [2, 3].

Section 3 starts with the unitization construction. Using this we will extend the result of section 2 to non unital commutative C*-algebras. We then go on with introducing positive elements and establishing some of their most important properties. The section ends with a proof for the existence of approximate units.

Section 4 starts with a general treatment of *-representations. We will work out dierent characterizations of irreducibility using von Neumann's bicommutant theorem. Afterwards we will introduce the notion of positive functionals in order to obtain representations via the Gelfand-Naimark-Segal construction (GNS con- struction). Ultimately, considering direct sums of GNS-representations, we prove the Gelfand-Naimark theorem which ensures the existence of an irreducible isomet- ric representation for every C*-algebra.

Most of the results and sketches of proofs are taken from [1]. Furthermore some results and proofs are taken from [4].

(3)

2. Basics

In this rst part we will review important denitions and facts. For proofs cf.

[2].

2.1. Operators on Hilbert spaces.

Denition 1. Given a Hilbert spaceH, we deneB(H)as the space of bounded linear operators onH.

Later on, we will have to deal with direct sums of Hilbert spaces and operators on them.

Denition 2. Let(Hi)i∈be a family of Hilbert spaces. The direct sum ⊕

i∈Ω

Hi is the space of tuples(ξi)i∈Ω, where P

i∈Ω

ik2<∞. With the inner product given by ((ξi)i∈Ω,(ηi)i∈Ω) =X

i∈Ω

i, ηi)

this space becomes a Hilbert space.

Given operators Ti ∈ B(Hi), i∈ Ωsuch that sup

i∈Ω

kTik <∞, one can consider their direct sum ⊕

i∈Ω

Ti dened by (⊕

i∈ΩTi)(ξi)i∈Ω = (Tiξi)i∈Ω.

i∈Ω

Ti is bounded with

i∈Ω

Ti

= sup

i∈Ω

kTik.

If all Hi are the same H, the direct sum is called the amplication of H by card(Ω). In that case, the amplicationT ∈B(⊕

i∈ΩH)of a T ∈B(H)is dened by

T= ⊕

i∈ΩTi.

Recall some of the most important topologies onB(H). For their denition we remember, that a topology is characterized by its convergent nets.

Denition 3. LetH be a Hilbert space

• The weak topology is dened byTi→T ⇐⇒(Tiξ, η)→(T ξ, η), ξ, η∈H.

• The strong topology is given byTi→T ⇐⇒ k(T−Ti)ξk →0, ξ∈H. Remark 4. The strong topology is induced by the family of semi-norms px: T 7→

kT xk.LetX be a vector space with a topology induced by a family of semi-norms (pi)i∈I. We recall, that for every x∈X a neighborhood basis is given by

{{y∈X: pi(x−y)< , i∈J}: J⊆Inite, >0}.

We will also need the Spectral Theorem for self-adjoint operators. For the proof and the denition of spectral measures cf. Section 7 in [2].

Theorem 5. Let H be a Hilbert space and A ∈ B(H) be a self-adjoint operator.

Let A be the σ-algebra of Borel sets on σ(A). Then there exists a unique spectral measureE on(σ(A),A, H), such that

A= ˆ

t dE(t).

(4)

Moreover for T∈B(H)

T A=AT ⇐⇒T E(∆) =E(∆)T, ∆∈ A.

The following denition will be useful.

Denition 6. LetB⊆B(H)andX ⊆H. Set

BX:={T x : T ∈B, x∈ X} Finally, we will state a standard argument as a lemma.

Lemma 7. Let H be a Hilbert space,S, T ∈B(H)andD⊆H dense. If (Sξ, η) = (T ξ, η) for allξ, η∈D,

thenS=T.

Proof. We have ((S −T)ξ, η) = 0 for all ξ, η ∈ D. By the density of D and continuity of the inner product, this holds forη ∈H. Thus(S−T)ξ∈H={0}

for allξ∈D. Again the density ofDyieldsS =T. 2.2. Banach algebras.

Denition 8. A Banach algebra is a Banach space(A,k.k)together with an asso- ciative and distributive multiplication·: A×A→A, such that

kx·yk ≤ kxkkyk for allx, y∈A.

It is called unital if there exists a multiplicative unit. This unit is unique and will be denoted by1. The set of invertible elements is denoted by

Inv(A) =

x∈A: ∃x−1∈A: xx−1=x−1x= 1 . Furthermore the spectrum of an elementx∈Ais given by

σ(x) ={λ∈C: x−λ1∈/ Inv(A)}. The spectral radius is given by

r(x) = max{|λ|: λ∈σ(x)}.

Remark 9. Note, that forx∈A, the multiplicative inversex−1 is unique.

The following properties concerning the spectrum hold true.

Theorem 10. LetA be a unital Banach algebra andx, y∈A. Then (1) σA(x)is nonempty and compact.

(2) r(x) = limn→∞kxnk1/n. In particular r(x)≤ kxk. (3) σA(xy)∪ {0}=σA(yx)∪ {0}

(4) If B is a Banach sub-algebra of A containing 0, then σA(x)⊆σB(x) and δ(σB(x)) ⊆ δ(σA(x)) where δ is the topological boundary. Equivalently, ρB(x)is open and closed as subset of ρA(x).

(5) σA(x−1) =

λ−1:λ∈σA(x)

Many proofs that technically assume the Banach algebra to be unital, can be ex- tended to the non unital case by the following construction.

Lemma 11. Let A be a non unital Banach algebra. Then A1 := A⊕C with coordinate wise addition, multiplication dened by(x;λ)(y;µ) = (xy+µx+λy;λµ) andk(x;λ)k1:=kxk+|λ|is a unital Banach algebra.

(5)

Elements of a Banach algebra Acan also be viewed as elements ofB(A), if we identify them with their respective multiplication operators.

Denition 12. LetAbe a Banach algebra. Forx∈A, we dene the multiplication operator Mx∈B(A)by

y7→xy.

Remark 13. The map x7→ Mx from A to B(A) is a homomorphism. Moreover, sincekxyk ≤ kxk kyk, we havekMxk ≤ kxk.

2.3. C*-Algebras.

Denition 14. A Banach*-algebra is a Banach algebraAwith a conjugate-linear and isometric involution∗:A→Asuch that

(xy)=yx for allx, y∈A.

If additionally the so called C*-axiom

kxxk=kxk2 for allx∈A holds,Ais called a C*-algebra.

Let A, B be Banach*-algebras. An algebra-homomorphism φ : A → B that satisesφ(x) =φ(x) for allx∈Ais called a *-homomorphism.

A sub-algebraS⊆A, that is *-closed, i.e. S⊆S, is called a *-sub-algebra.

Example 15. The following examples are fundamental in the sense that all C*- algebras, as we will see, can be represented as a *-sub-algebra of one of them.

(1) (Unital commutative C*-algebras) Let C(X) be the set of all continuous complex-valued functions on a compact topological space X. Equipped with point-wise operations, the point-wise complex-conjugate and the norm given bykfk= sup{f(x) : x∈X}it is a C*-algebra.

(2) (Non unital commutative C*-algebras) We consider the space C0(X) ={f ∈C(X) : ∀ >0∃K compact: f(X\K)⊆U(0)}

for a locally compact spaceX. With the same structure as above one again obtains a C*-algebra.

(3) (Non commutative C*-algebras) Consider the spaceB(H)of bounded linear operators on a Hilbert spaceH. With the operation which maps an operator to its adjoint, the usual algebraic operations - multiplication being the composition of operators - and the operator-norm,B(H)is a C*-algebra.

In analogy to operators we can dene normal, self-adjoint and unitary elements for C*-algebras.

Denition 16. For a *-algebraA,x∈Ais called

• normal ifxx=xx

• self-adjoint ifx=x

• unitary ifx=x−1for unitalA.

Remark 17. The set of self-adjoint elements is denoted byAsa. We recall the following facts.

Theorem 18. LetA be a unital C*-algebra andx∈A.

(1) Ifxis unitary, thenσA(x)⊆Twhere Tdenotes the unit circle.

(6)

(2) Ifxis self-adjoint thenσA(x)⊆R.

Lemma 19. Let B be a unital C*-sub-algebra of a unital C*-algebraA (with the same unit), thenσB(x) =σA(x)for allx∈B.

A main characteristic of C*-algebras is, that algebraic (spectrum) properties are strongly linked to topological (norm) properties. This connection roots in the in- nocuous looking C*-axiom.

Lemma 20. For a unital C*-algebra, letx∈Abe normal. Thenkxk=r(x)where r(x)is the spectral radius. Furthermore,kxk2=r(xx)for arbitraryx∈A.

This also results in additional structure for *-homomorphisms.

Lemma 21. Letϕ:A→Bbe a *-homomorphism between unital C*-algebrasA, B, such that ϕ(1A) = 1B. Then ϕis automatically bounded with kϕk ≤1. Moreover, σ(ϕ(x))⊆σ(x)x∈A.

Remark 22. The previous lemma also holds true for the weaker assumption of A being a Banach *-algebra.

2.4. Unital Commutative C*-algebras. For the proofs in this section, the reader is referred to Section 1 in [3].

Denition 23. Let A be a C*-algebra. A multiplicative functional φ:A→C is a linear functional additionally satisfyingφ(ab) =φ(a)φ(b). The Gelfand space or spectrum ofAis dened by

Aˆ:={φ:A→C|φ is multiplicative; φ6= 0}.

Lemma 24. Let φ 6= 0 be a multiplicative functional on a C*-algebra A. Then kφk= 1. IfA is unital, then φ(1) = 1.

We recall that thew-topology on any subsetY ⊆A of the topological dual of a Banach spaceA, is the initial topology with respect to the maps

ι(a) : Y →C; φ7→φ(a),

where a ∈ A. We denote it by σ(Y, ι(A)). The following Banach-Alaoglu type theorem for unital C*-algebras holds true.

Theorem 25. Let A be a unital C*-algebra. Its Gelfand space Aˆ is a compact subspace of the topological dual spaceA with respect to thew-topology.

Remark 26. From now on, if not otherwise stated explicitly, we assume Aˆ to be endowed with thew-topology.

The main result of this section is the following representation theorem.

Theorem 27. LetA be a commutative unital C*-algebra. The map ι:A→C( ˆA) dened byι(a) = (φ7→φ(a))is an isometric isomorphism.

Denition 28. LetA be a unital C*-algebra. For a normalx∈Adene C(x) :=c.l.s.

xj(x)k:j, k∈N0

wherex0:= 1.

More generally for commutingx1, x2, . . . , xn, each of which is normal, we dene C(x1, x2, . . . , xn)to be the closed linear span of the products of the powers of the xk andxk.

(7)

Remark 29. It is easy to see, thatC(x)is a commutative C*-sub-algebra ofA. Indeed, it is the smallest one containingxand1.

As a corollary of Theorem (27) we obtain the continuous functional calculus for normal elements of unital C*-algebras.

Corollary 30. Let A be a unital C*-algebra. For any normal element x ∈ A, C(x) is isometrically isomorphic toC(σ(x)). The isomorphism Φmaps xto the identity id : σ(x)→C; t 7→t. For f ∈C(σ(x))we dene f(x) := Φ−1(f). The following properties hold

(1) σ(f(x)) =f(σ(x)).

(2) Iff ∈C(σ(x)),g∈C(σ(f(x))), theng◦f ∈C(σ(x))andg◦f(x) =g(f(x)). Proof. The proofs found in Section 1 in [3] may be applied to the present situation.

Remark 31. This functional calculus may be used to prove the spectral theorem for normal operators.

Lemma 32. Let A, B be unital C*-algebras, ϕ: A → B a *-homomorphism be- tween them, such thatϕ(1A) = 1B andx∈Anormal. Thenϕ(f(x)) =f(ϕ(x))for any f ∈C(σ(x)).

Proof. We consider the unital commutative sub-algebras C(x)⊆A, C(ϕ(x))⊆ B and the induced homomorphism between them. If f(z) = Pn

j,k=0λj,kzjk is a polynomial, the claim is obvious. By Stone-Weierstrass these polynomials are dense in C(σ(x)). From Lemma (21) we know that σ(ϕ(x) ⊆ σ(x). Denote by

|σ(ϕ(x)) : C(σ(x)) → C(σ(ϕ(x)) the restriction mapping and by Φxϕ(x)) the mapping from Corollary (30) applied tox(ϕ(x)).We just showed, that

ϕ◦Φ−1x (f) = Φ−1ϕ(x)◦ |σ(ϕ(x))(f)

for anyf ∈C[z,z]¯ viewed as an element ofC(σ(x)).According to the continuity of all involved mappings and due to the density of C[z,z]¯ we see that the above equation holds true for allf ∈C(σ(x)), i.e. ϕ(f(x)) =f(ϕ(x)). As a rst application of the functional calculus, we will strengthen the statement from Lemma (21).

Lemma 33. Let φ:A→B be a *-homomorphism such that φ(1A) = 1B between unital C*-algebrasA, B. Ifφis injective it is an isometry.

Proof. We already know, thatkφ(x)k ≤ kxkfor allx∈A, cf. Lemma (21). Let us suppose there is anx∈A, such that this inequality is strict. Then also

r:=kφ(xx)k<kxxk=:s.

We now consider the C*-algebraC(xx)(note, thatxxis normal) and the induced injective *-homomorphism φ : C(xx) → B. It is indeed injective since for x∈ A∩C(xx)we haveφ(x+λ1) =φ(x) +λ1whereφ(x)and1are obviously linearly independent. Consider the function f :=g|σ(xx), where g ∈C([−s, s]) vanishing on[−s, r]withg(s) = 1. By the functional calculus we have

0 =f(φ(xx)) =φ(f(xx)).

Sincef does not vanish onσ(xx), the equality above contradicts the injectivity

ofφ.

(8)

3. Further Structure

In the followingA, B, ...will always denote a C*-algebra unless otherwise stated.

Moreover, the word functional always means linear functional.

3.1. Unitization. A C*-algebra need not be unital. However, there is a canonical way to construct an enveloping unital C*-algebra.

Remark 34. If A is a C*-algebra, the map x 7→ Mx (cf. Denition (12) ) is an isometric homomorphism. OnlykMxk ≥ kxk is left to proof. Indeed

kMxk ≥

Mx x kxk

= 1

kxkkxxk=kxk.

Theorem 35. LetAbe a non unital C*-algebra. ThenA1:=A⊕Cwith coordinate wise addition, (x;λ)(y;µ) = (xy+µx+λy;λµ) , k(x;λ)k1 = sup

kyk=1

kxy+λyk and (x;λ)= (x;λ)is a unital C*-algebra. The unit is given by (0; 1).

Proof. k.k1 is the operator norm on A1 if its elements are viewed as sums of left multiplication operators and multiples of the identity onA, i.e. (x, λ)'Mx+λI. Thusk.k1 actually is a norm and it also satiseskxyk1≤ kxk1kyk1.

To show completeness we rst notice that equipped with the normk(x;λ)k2:=

kxk+|λ|and the algebraic rules given above, A1 is a Banach algebra; (cf Lemma (11)).

The next step is to show, thatk.k1 and k.k2 are equivalent. Sincekxy+λyk ≤ kxkkyk+|λ|kyk=k(x;λ)k2for every kyk= 1,

k(x;λ)k1≤ k(x;λ)k2.

Suppose there is no n ∈ Nwith k(x;λ)k2 ≤nk(x;λ)k1 for all (x;λ) ∈ A1. Then there are sequencesxn, λn withk(xnn)k2≥nkxny+λnykfor allkyk= 1. Take y=xn/kxnk to see, thatλn6= 0 forn >1. Settingzn:=xnn one gets

(3.1) kznk+ 1≥nkzny+yk for all kyk= 1.

(zn)n∈

Ncannot be unbounded since for everyn∈Nwe may takey:=zn/kznk, to obtain

kznk+ 1≥nkzny+yk ≥nkznk −n.

Hence

n+ 1

n−1 ≥ kznk, n >1.

So by (3.1) for alln, m≥n()andkyk= 1we getkzny+yk ≤ε/2and therefore k(zn−zm)yk ≤ kzny+yk+kzmy+yk ≤.

By the previous remark (zn) is a Cauchy sequence. Thus it converges to a limit z∈A satisfyingzy=−y for ally. But this is impossible becauseAis non unital.

Sok...k1 is a norm equivalent tok...k2 showing that(A1;k...k1)is complete.

Next we prove that the-operation is an isometry and that the C*-axiom holds true. In fact

k(x;λ)k21= sup

kyk=1

kxy+λyk2= sup

kyk=1

k(xy+λy)(xy+λy)k

= sup

kyk=1

kyxxy+λyxy+ ¯λyxy+|λ|2yyk ≤

(9)

≤ sup

kyk=1

kxxy+λxy+ ¯λxy+|λ|2yk=k(xx+λx+ ¯λx;|λ|2)k1

=k(x;λ)(x;λ)k1≤ k(x;λ)k1k(x;λ)k1.

We get k(x;λ)k1 ≤ k(x;λ)k1 and by symmetry equality. In turn, we see, that the above inequality is an equality, proving the C*-axiom. The remaining properties are easily veried by a straight forward calculation.

Remark 36. By Remark (34) we have thatk.k1|A=k.kfor a non unitalA. There- fore, A can be viewed as the subset A⊕ {0} of A1 by the isometric embedding a7→(a; 0). As such it is a closed ideal withA1/A∼=C.

Denition 37. LetAbe a unital C*-algebra. IfAis unital, then we set. Otherwise we set A1 := A⊕C and provide this space with the C*-algebra structure from Theorem (35). Moreover, for a normalx∈Awe dene

C(x) :=c.l.s.

xj(x)k:j, k∈N0 ⊆A1

where x0 := 1.More generally for commutingx1, x2, . . . , xn, each of which is nor- mal, we deneC(x1, x2, . . . , xn)to be the closed linear span of the products of the powers of thexk andxk.

Lemma 38. Let X be a non compact but locally compact Hausdor space. If ac(X)denotes the one-point (Alexandro) compactication of X, then C0(X)1∼= C(ac(X))i.e. there exists an isometric *-isomorphism between those C*-algebras.

Proof. We show, that the mapΦ :C0(X)1→C(ac(X))dened by Φ((f;λ))(x) =

( f(x) +λ x∈X

λ x=∞

is an isometric *-isomorphism.

It is easy to see, that the map is well-dened and a *-isomorphism. By Lemma (33) it is also isometric .

Alternatively we can show isometry directly. In order to do so, note thatf ∈ C0(X) takes its maximum at a point y ∈ X. By denition we get k(f;λ)k1 = sup{kf+λgk:g∈C0(X), kgk= 1}=f(y) +λ. Since,kΦ((f;λ))k=f(y) +λ,Φ

is an isometry.

Remark 39. LetA, B be C*-algebras and ϕ:A→B a *-homomorphism. IfA is unital, then by the continuity of the product ϕ(1A)is a unit of the *-subalgebra ϕ(A). Nevertheless it can happen, thatB is not unital or that 1B6=ϕ(1A).

An example for1B 6=ϕ(1A)is easily constructed:

LetM(n)denote the C*-algebra ofn-dimensional matrices (cf. Example (15.3)).

Then the embedding

ι:M(m)→M(n), m < n given by

A7→

A 0 0 0

is a *-homomorphism between this C*-algebras. Obviously, we haveι(En)6=Em.

(10)

Remark 40. LetA, B be C*-algebras and ϕ:A→B a *-homomorphism. IfA is non unital, then it is easy to check, that(x;λ)7→φ(x) +λ1is a *-homomorphism fromA1toB1extending ϕ. This extension will also be denoted byϕ.

Ifϕ(A)is dense inB, thenϕ:A1→B1 is the unique extension ofϕ:A→B to a homomorphism. Assume, thatψ:A1→B1is another extension ofϕ:A→B. If B=B1, i.e. Bis unital, then it follows from the previous remark, that1B=ψ(1A1) and by linearityψ=ϕonA1. IfB is not unital, thenψ(1A1)∈/B, since otherwise ψ(1A1) would be a unit in B, which contradicts our assumption. Consequently, ψ(A1)is dense inB1. As before, we derive1B1 =ψ(1A1)and in turnψ=ϕonA1. Lemma 41. Any *-homomorphism ϕ : A → B between C*-algebras A, B is a contraction. Ifϕis injective, then it is isometric.

Proof. ReplacingB by the closure ofϕ(A), we can assume ϕto have dense range.

We saw in the previous remark, thatϕthen admits a unique extensionϕ:A1→B1

which according to Lemma (21) is a contraction.

Ifϕis injective andAis not unital, then1B1∈/ϕ(A), since otherwise injectivity would give the existence of a unit in A. Thus, also ϕ : A1 → B1 is injective.

According to Lemma (33)ϕis isometric.

Corollary 42. IfAis a non-unital C*-algebra andB a unital C*-Algebra contain- ingA, then A+c.l.s.{1B} is isometrically isomorphic toA1.

Proof. According to Remark (40) the inclusion mapping A → B extends to an injective *-homomorphism. By the previous lemma this extension is an isometric

isomorphism.

Remark 43. For a C*-algebraAandx∈A, letσ(x) =σA(x)be the spectrum ofx, when xis considered as an element of A1. By the previous corollary and Lemma (19) we haveσA(x) =σB(x)for any C*-algebraB⊆A.

3.2. Commutative C*-algebras. In this section we will expand the results of Section 2.3. to the non unital case. We start with examining the Gelfand space of a unitization of a non unital C*-algebra A.

Remark 44. By Lemma (24) we haveφ(1) = 1for allφ∈Aˆ1(cf. Def (23)). Thus by Remark (40), we can uniquely extend anyφ∈Aˆ to a multiplicative functional onA1by settingφ(1) = 1. By default we will identifyφwith its unique extension.

Moreover, since it also holds, thatφ|Ais a multiplicative functional onAfor any φ∈Aˆ1, we getAˆ1= ˆA∪ {φ1}whereφ1|A= 0andφ1(1) = 1.

Indeed an even stronger proposition is true.

Lemma 45. Let A be a non unital C*-algebra. Aˆ1 is homeomorphic to the one point compactication ac( ˆA)of Aˆ.

Proof. We use the notation introduced previous to Theorem (25). By the remark above,( ˆA, σ( ˆA, ι(A)))is homeomorphic to( ˆA1\ {φ1}, σ( ˆA1\ {φ1}, ι(A1))). To see this let Ψ : ˆA →Aˆ1\ {φ1} map φ∈Aˆ to its unique extension. Obviously this is a bijection. For alla∈A one hasι(a;λ)◦Ψ = ιa+λ andιa◦Ψ−1(a;0) . The functions on the right side of the equation are continuous. Thus by the denition of the respective initial topologies,Ψis bi-continuous.

(11)

By Theorem (25)Aˆ1 is compact. Furthermore Aˆ1\ {φ1} is open and dense in ( ˆA1, σ( ˆA1, ι(A1))). So using the uniqueness of the one point compactication we

get, thatAˆ1is homeomorphic toac( ˆA).

Corollary 46. ( ˆA,σ( ˆA, ι(A))) is locally-compact (the existence of a unit is not required).

Proof. We recall, that a space is locally compact if its one-point compactication is Hausdor. Indeed, by the above lemma, the one point compactication is home-

omorphic to the Hausdor spaceAˆ1.

We now have everything in order, to extend Theorem (27) to the non unital case.

Corollary 47. LetA be a commutative C*-algebra. Then A is isometrically iso- morphic toC0( ˆA).

Proof. We already know, thatA1 is isometrically isomorphic toC( ˆA1), so we can assume, that A is non-unital. By Lemma (45) Aˆ1 is homeomorphic to ac( ˆA). Let Φ : ˆA1 → ac( ˆA) be the corresponding a homeomorphism. Then the map Υ(f) = f ◦Φ from C( ˆA1) to C(ac( ˆA)) is an isometric isomorphism. We now use Lemma (45) to see, thatC(ac( ˆA))is also isometrically isomorphic to C0( ˆA)1. Putting it all together on gets an isometric isomorphismΩ :A1→C0( ˆA)1. Taking a closer look on this isomorphism one sees, thatΩ(A) =C0( ˆA), where we consider C0( ˆA)as a subset ofC0( ˆA)1 by Remark (36).

3.3. positive Elements. LetAbe a C*-algebra for the remainder of this section.

Denition 48. A self-adjoint xis called positive if σ(x)⊆[0,∞). The set of all positive elements is denotedA+. We also writex≥0to indicate, thatxis positive.

Remark 49. Sinceσ(f) =f(X)forf ∈C(X)and a compact spaceX, the positive elements ofC(X)are exactly the non negative functions. By Lemma (45) the same is true forC0(X).

It is useful to know, that every element can be written as a linear combination of four positive elements.

Lemma 50. Lety∈Asa andx∈Athen

(1) x = Re(x) +iIm(x) where Re(x) = x+x2 and Im(x) = x−x2i are self- adjoint.

(2) Fory∈Asaone has the decompositiony=y+−ywherey+, y∈A+are unique satisfyingy+y= 0. We also haveσ(y)∩(−∞,0) =σ(y)∩(−∞,0) andσ(y+)∩(0,∞) =σ(y)∩(0,∞).

Proof. The rst assertion is immediately seen by a straightforward calculation.

For (2) consider the representation of the (commutative) C*-algebra C(y). It is easy to see, that y+, y represented by f(t) := max (t,0), g(t) := −min (t,0) respectively, satisfy the conditions. The claim about the spectra follows from the

previous theorem.

From the functional calculus in the previous section one derives some basic facts about positive elements.

Lemma 51. Forx∈A, the following assertions hold true:

(12)

(1) Ifxis normal thenxx≥0.

(2) Ifx∈Asa andkxk ≤2, thenx≥0⇐⇒ k1−xk ≤1(inA1).

(3) Ifx≥0 then there exists a square-rootx1/2≥0. (4) (x;λ)≥0 inA1if and only ifx∈Asa andλ≥ kxk.

Proof. In each case it is possible to consider the C*-algebra C(x) instead of A. Therefore, xcan be represented by some f ∈C(σ(x)). The rst three properties then follow immediately using that fact. Concerning (4), note that σ((x;λ)) = σ(x) +λ. Then

σ((x;λ))∩(−∞,0) =σ((x;λ))∩(−∞,0) = (σ(x) +λ)∩(−∞,0)

yields the fourth property.

We continue with some further properties. Note, that (2) is a considerable improve- ment of (1) in the previous lemma.

Theorem 52. Letx∈A.

(1) Ifx, y≥0thenx+y≥0, i.e. A+ is a closed cone.

(2) xx≥0.

(3) xyx= (y1/2x)(y1/2x)≥0 fory≥0. Proof. (1) By Lemma (51) we have,

A+∩K1(0) =Asa∩K1(0)∩ {x:k1−xk ≤1}

with K1(0) being the closed unit ball in A. SinceAsa is closed by the continuity of , this is a closed and convex set. For x, y∈ A+, there is a C >0 such, that

x

C,Cy ∈A+∩K1(0). By convexity 1

2C(x+y)∈A+∩K1(0).

Since positive multiples of positive elements are positive we havex+y∈A+. (2) To prove the second property, according to Lemma (50) we decomposexx= c+−c and deneh:=xc. A calculation yields

−hh=c3≥0.

Using Lemma (51),(1) along with (1) we obtain

hh= (hh+hh) + (−hh) = 2Re(h)2+ 2Im(h)2+ (c3)≥0.

Sincehh≥0⇐⇒hh≥0 by Theorem (10),(3), it follows thatc = 0.

(3) immediately follows from (2).

Remark 53. By property (2) and the existence of square roots we get thatx∈Ais positive if and only if there existsy∈Asuch that x=yy. This characterization does not hold in general Banach *-algebras. The reason for this is that one does not have the functional calculus which is used to obtain the decomposition xx= c+−c.

Denition 54. Forx, y∈Aone writesx≥y ifx−y≥0.

(13)

Remark 55. This relation denes a partial order onAsatisfyingx≤y⇒x+z≤ y+z for allz∈A. To assert transitivity letx≤yandy≤z. Then

z−x= (z−y) + (y−x)≥0

because the sum of positive elements is positive by the previous theorem. We getx≤z.

Lemma 56. Ify≤z, then xyx≤xzxand if y∈A+, then0≤xyx≤ kykxx. Proof. By (3) of the previous theorem, we have x(z−y)x ≥ 0 . Hence we get xyx≤xzx. From Theorem (10),(2) we obtainy≤ kyk1. The second inequality immediately follows from this fact and the rst part of the present assertion.

The next lemma and its corollary will be used in the next section, to proof the existence of approximate units.

Lemma 57. If0≤x≤y andx∈Inv(A) theny∈Inv(A) and0≤y−1≤x−1. Proof. Ifx∈Inv(A)∩A+ then ε1 ≤xfor some >0 as one sees by identifying xwith f ∈C(σ(x)). Therefore, also 1 ≤y. Identifyingy with a functiong, we see, that y is invertible. If x andy commute, we may restrict our considerations toC(x, y)and use representation theory from Theorem (27) for commutative C*- algebras to get the inequaltiy. Indeed, xandy are both represented by functions f and g, on the Gelfand space of C(x, y), respectively, and we have 0 ≤f ≤g. Thus0≤g−1≤f−1and nally0≤y−1≤x−1. For the general case note, that by Lemma (56)

y−1/2xy−1/2≤y−1/2yy−1/2= 1.

From the special case we get

1≤(y−1/2xy−1/2)−1=y1/2x−1y1/2. Finally,

y−1=y−1/21y−1/2≤y−1/2y1/2x−1y1/2y−1/2=x−1.

Corollary 58. If 0≤x≤y, thenkxk ≤ kyk.

Proof. Let λ > r(y), then 0 ≤ λ−y ≤ λ−x . By the previous lemma, since λ−y∈Inv(A), λ−xis also invertible. Therefore,kxk=r(x)≤r(y) =kyk. 3.4. Approximate Units.

Denition 59. An approximate unit for a C*-algebraAis a net(hλ)λ∈Λof positive elements inAwithkhλk ≤1 andhλx→xfor allx∈A.

There always exists an approximate unit for a C*-algebra. Indeed for any dense two-sided idealI⊆Athere is an approximate unit contained inI.

Theorem 60. IfI⊆Ais a dense two-sided Ideal one denesΛI :={x∈I+:kxk<1}

with I+ :=A+∩I. Then there exists a direction onΛI, given by the partial order

≤, such that it is an approximate unit.

(14)

Proof. Step 1: We start showing, that ≤is a direction on ΛI. To do so, we will prove, that the mapα: x7→(1−x)−1−1denes an order-isomorphism from ΛI

ontoI+, where the order is given by≤.

Indeed, forx∈ΛI,α(x)is well dened and belongs toI+, since

(1−x)−1−1 =

X

k=1

xk =x

(X

k=0

xk)∈I+

bykxk<1. By Lemma (57)αis order preserving.

Ify∈I+then, (1−(1 +y)−1)(1 +y) =y.

The invertibility of(1 +y)(cf. Lemma (57)) yields 1−(1 +y)−1=y(1 +y)−1∈I+.

Lety be represented by the continuous function g on the Gelfand space ofC(y). Then y(1 +y)−1 is represented by 1+gg . The compactness of the Gelfand space yieldsky(1 +y)−1k<1. Therefore,

1−(1 +y)−1∈ΛI.

Thus the mappingβ : x7→1−(1 +x)−1mapsI+ontoΛI and obviously satises α◦β=idI+,β◦α=idΛI. Hence, αis a bijection with the inverse mappingβ.

AsI+is a directed set (x, y∈I+⇒x, y≤x+y), the same is true forΛI. Step 2: In order to prove, thatΛI is an approximate unit, we have to show, that the net(x(1−y)x)y∈Λ

I converges to0 for every x∈ΛA. Indeed from(1−y)2= (1−y)12(1−y)(1−y)12 ≤(1−y)we get

k(1−y)xk2=

x(1−y)2x

≤ kx(1−y)xk.

Hence, from the fact, that by Lemma (50) everyx∈A can be written as a linear combination of elements fromΛA={x∈A+:kxk<1}this suces.

We rst notice thatΛI is dense in ΛA. Indeed, let y ∈I be a sucently good approximation ofx12 forx∈ΛA, then by

kx−yyk ≤ kx12(x12 −y)k+k(x12 −y)yk

As a consequence without loss of generality we may assume that I = A. We now xx∈ΛA and viewxas a function on the Gelfandspace of C(x). For every > 0, that is suciently small, there exists a point µ ∈ n

λ∈Cˆ(x) :f(λ)≥o independent of, wheref obtains its maximum. We see, that for those >0 one nds ann∈Nsuch that

kx(1−xn1)xk ≤

f(1−fn1)f

≤maxn

f(µ)(1−n1)f(µ), 2o

≤. By Lemma (56), we see thatx(1−y)xis decreasing. Therefore,(x(1−y)x)y∈Λ converges to0 and the proof is completed. I

we get thatyy∈ΛI approximates x.

The next lemma establishes some properties of approximate units.

Lemma 61. If(hλ)λ∈Λis an approximate unit in Athen (1) xhλ→xandhλxhλ→xfor allx∈A.

(2) (hαλ)λ∈Λ is an approximate unit for anyα >0.

(15)

Proof. 1) Sincehλ is positivexhλ→xfollows fromhλx→x. By decomposition one can assume thatx≥0. Then

hλxhλ= (hλx12)(x12hλ)→x12x12 =x.

2) One has

kh2λx−xk ≤ kh2λx−hλxhλk+khλxhλ−xk=khλk khλx−xhλk+khλxhλ−xk →0.

By induction

kh2λnx−xk →0

for anyn∈N. Hence, there exits a λ0, such that for allλ≥λ0, we have kxh2λnx−xxk ≤ε.

Ifα >0, choosenwith2n−1≥α. Then

xh2λnx≤xhλ x≤xhαλx≤xx.

Combining these inequalities, by Corollary (58), we obtain kxhαλx−xhλ xk ≤ε, kxx−xhαλxk ≤ε.

Therefore,

kx−hαλxk2=k(x−xhαλ)(x−hαλx)k=kxx−2xhαλx+xhλ xk

≤ kxx−xhαλxk+kxhαλx−xhλ xk ≤2ε.

Remark 62. For the approximate unit constructed in Theorem (60) there is an easier way to see (2). Indeed, if we examine the last part of the proof of Theorem (60), we see, that there also exists ann∈N, such thatkx(1−xα/n)xk ≤.

(16)

4. Representation Theory of C*-Algebras 4.1. Representations.

Denition 63. A pair(H, π)is called a representation of a Banach *-algebraAif π: A→ B(H) is a *-homomorphism into the space of bounded operators on the Hilbert spaceH.

Remark 64. Sometimes we will call justπa representation if it is clear what Hilbert space is meant.

We introduce some notions for representations.

Denition 65. Let(H, π)be a representation of a Banach *-algebraA.

• It is called faithful ifπis injective.

• It is called irreducible if there are no non trivial closed invariant subspaces U, i.e. π(A)(U)⊆U ⇒U ={0} ∨U =H.

• LetB ⊆B(H) be a *-sub-algebra (i.e. a sub-algebra satisfyingB⊆B).

Consider the largest subspace N ⊆ H such that B|N = 0 ; i.e. N = T

A∈BkerA. Its orthogonal complement X :=N is called the essential subspace ofB. If X =H, we say thatB acts nondegenerately onH. The representationπis called nondegenerate ifπ(A)⊆B(H)acts nondegener- ately onH.

• ξ∈H is called a cyclic vector ifπ(A)ξ=H. If there exists a cyclic vector ξ, we call(H, π, ξ)a cyclic representation.

Remark 66. If a representationπis irreducible, then it is also nondegenerate.

Irreducibility is formulated in terms of closed invariant subspaces. Since SU ⊆ SU for all bounded operators S, we haveπ(A)U ⊆U ⇒π(A)U ⊆U. Hence, the only invariant non trivial subspaces of an irreducible representation must be dense inH.

Lemma 67. B ⊆B(H)acts nondegenerately if and only ifBH=H.

Proof. IfBH =:X 6=H , thenN :=X 6={0}. SinceX is invariant and B is a

*-sub-algebra,N also is. To see this takex∈N, then (Bx, X) = (x, BX) = (x, BX)

where BX ⊆X, becauseX is invariant. It follows that(Bx, X) ={0}. Hence BN ⊆N. asBN ⊆BH=X⊥N,B acts degenerately. Now supposeB|N = 0on someN 6={0}. Again, becauseN is invariantX :=N6=H also is. Forx∈H we have(Bx, N) = (x, BN) ={0}.Hence,BH⊆X and furtherBH⊆X =X.

Before we give some characterizations of irreducibility we have to prove the von- Neumann bicommutant theorem.

Denition 68. For a Hilbert spaceH and A ⊆B(H)we dene the commutant A0 :={T ∈B(H) :T S=ST for allS∈A}.

Remark 69. ObviouslyA0 ⊆B(H)is a sub-algebra. If A is closed, than A0 also is. Indeed

ab= (ba)= (ab)=ba fora∈A0 and allb∈A.

Moreover, we haveA⊆A00. Therefore,A0 ⊆A000. SinceA000 commutes withA00, and hence also withA, we getA000 ⊆A0 and nally A0 =A000.

(17)

Lemma 70. LetA⊆B(H). Then the commutantA0is closed in the weak topology.

In particular, it is strongly closed.

Proof. Let(Ti)i∈I ∈A0 be a net, with(Tiξ, η)→(T ξ, η)for allξ, η∈H. We need to prove, thatT ∈A0. For an arbitraryS∈A, we get

(T Sξ, η) = lim

i∈I(TiSξ, η) = lim

i∈I(STiξ, η) = lim

i∈I(Tiξ, Sη) = (T ξ, Sη) = (ST ξ, η).

By Lemma (7)T S=ST.

Theorem 71. (Bicommutant) LetH be a Hilbert space. If B ⊆B(H)is a *-sub- algebra that acts nondegenerately, thenB is strongly dense inB00.

Proof. LetT00∈A00. We have to show, that every strong neighborhood (cf. Remark (4))

N,x1,...,xn(T00) =n

T ∈B(H) :

(T00−T)ξj

< , j= 1, . . . , no contains an element ofB. Fixingξ1, ξ2, . . . , ξn∈H and >0a T ∈B with

maxn

(T00−T)ξj

: j= 1, . . . , no

<

has to be constructed. Therefore, we consider the amplications (cf. Denition (2)) H:= ⊕

j=1,...,nH, S:= ⊕

j=1,...,nS

forS∈B(H). Moreover, we deneB ={S:S∈B}. ObviouslyB is a *-sub- algebra ofB(H)that acts nondegeneratly. Now it suces to show, that there is aT ∈B such that

(T−(T00))ξ <

whereξ= (ξ1, . . . , ξn).

For this considerX =Bξ. Since bothX andX are invariant underB, we havePX ∈(B)0 for the orthogonal projectionPX ontoX. Since(T00) commutes with PX, we get (T00)X ⊆X. If we can show that ξ∈X we are done, for then (T00)ξ∈X . For allS∈B we have

S[(1−PX)ξ] = (1−PX)[Sξ] = 0

sinceSξ∈X. BecauseB acts nondegeneratly we may conclude(1−PX)ξ= 0,

i.e. ξ∈X.

We will see, that we can reduce our attention to projections if we want to know whether someT ∈B(H)belongs to a certain commutant.

Lemma 72. LetB⊆B(H)be a *-sub-algebra andT ∈B(H). ThenT ∈B0 if and only if the ranges of all spectral projections of Re(T) and Im(T) are B-invariant subspaces.

Proof. ForT ∈B0 we haveRe(T), Im(T)∈B0 as they are linear combinations of T, T. It follows from Theorem (5) , that their spectral projections E(∆) are in B0. Therefore, their ranges are invariant subspaces. Conversely, if the ranges of all spectral projections are invariant subspaces, then all projections are in B0. Again by Theorem (5) we getRe(T), Im(T)∈B0. Hence,T ∈B0. Now we have all that we need to prove the following characterizations of irre- ducible representations.

(18)

Theorem 73. Let (π, H) be a representation of a C*-algebra A. The following assertions are equivalent:

(1) πis irreducible.

(2) π(A)0 =C1. (3) π(A)00=B(H).

(4) π(A)is strongly dense inB(H).

(5) Ifdim(H)6= 1 , then every06=ξ∈H is a cyclic vector forπ.

Proof. Lets start with 1) ⇔ 2). Suppose π is irreducible. If T ∈ π(A)0 , then by the previous lemma and the denition of irreducibility Re(T) and Im(T) are multiples of the identity. ThusT is a multiple of the identity. On the other hand ifπ(A)0 =C1, then again by the previous lemma the only invariant subspaces are trivial.

2) ⇒ 3) is trivial. For 3) ⇒ 2) note, that B(H) has no nontrivial invariant subspaces. By the previous lemmaπ0(A) =π(A)000 =B(H)0 =C1.

4)⇒3) is immediate by Lemma (70).

3)⇒4) Since 3) implies 1), this follows by the commutant theorem.

We will now prove 1)⇔ 5). Let πbe irreducible. Then for ξ∈ H, the closed invariant subspace π(A)ξ must be trivial. If π(A)ξ = {0}, then Cξ is a closed invariant subspace and thus the whole space, sodim(H) = 1.

For the other direction, consider a nonzero closed invariant subspaceU ⊆H. If dim(H) = 1, then triviallyU =H. If on the other hand everyξis a cyclic vector,

chooseξ∈U. ThenH=π(A)ξ⊆U.

Denition 74. LetAbe a C*-algebra and(π1, H1),(π2, H2)representations. We callT ∈B(H1, H2)an intertwining operator if

T π1(x) =π2(x)T

for everyx∈A. The set of all intertwining operators is denotedR(π1, π2). IfU ∈ R(π1, π2)is unitary(π1, H1),(π2, H2)are called unitarily equivalent. If in addition, there exist cyclic vectorsξ1, ξ2withU ξ12, then(π1, H1, ξ1),(π2, H2, ξ2) are called unitarily equivalent as cyclic representations.

4.2. Positive Functionals, States. In order to describe commutative C*-algebras we used multiplicative functionals. This approach fails in general, since a multiplica- tive functional cannot distinguish betweenxyandyx for non commuting elements x, y∈A. It turns out, that in order to describe a non-commutative C*-algebra one has to introduce the more general notion of states.

Denition 75. LetAbe a C*-algebra. A functionalωis called positive ifω(x)≥0 for allx∈A+. If it is normalized, i.e. kωk= 1, then it is called a state. The set of states onAis denoted by S(A).

Example 76. By Lemma (24) any multiplicative functional on a commutative C*-algebra is a state.

Remark 77. For general Banach *-algebras a functional is called positive ifω(xx)≥ 0 for allx ∈ A. Indeed this dening property is the one that is needed in what follows (see Lemma (81)). However by Remark (53) we see that this denition is equivalent to the one given above in the case of C*-algebras. Since we do not have this equivalence for general Banach *-algebras, a lot of the following proofs will not

(19)

work in that case. Nevertheless the majority of these results hold in more general settings as well, though the proofs are more elaborate.

Let us start by proving that every positive functional is automatically bounded.

Lemma 78. Let ω be a positive functional, then there exists C > 0 such that kωk ≤C.

It is an important fact, that any positive linear functional gives rise to a pre-inner product.

Proof. After decomposition (cf. Lemma(50)) it suces to show that|ω(x)| ≤Ckxk for allx∈A+and aC >0. Suppose there is no suchC. Then for everyn∈Nthere is xn ∈A+ with kxnk = 1 and ω(xn)≥4n. We now consider x:=P

n=12−nxn. This is a well dened element ofAsatisfyingω(x)≥2−nω(xn)≥2n for anyn∈N,

which cannot be true.

Denition 79. (x, y)ω:=ω(yx)for any positive functionalω onA. Lemma 80. Letω be a positive functional. Then ω(x) =ω(x). Proof. Ifx∈Asa, thenx=x+−x withx+, x ∈A+. We get

ω(x) =ω(x+−x) =ω(x+)−ω(x)∈R. Forx∈A,x=a+ibwitha, b∈Asa, we have

ω(x) =ω(a−ib) =ω(a)−iω(b) =ω(x)

sinceω(a), ω(b)∈R.

Lemma 81. (., .)ω is a pre-inner product onA.

Proof. (x, y)ω = ω(yx) = ω(xy) = (y, x)ω by Lemma (80), and (x, x)ω = ω(xx)≥0 by Theorem (52). The sesquiliniarity is obvious.

Since we have shown that (., .)ω is a pre-inner product, one gets the Cauchy- Schwarz inequality.

Fact 82. |ω(yx)|2≤ω(xx)ω(yy)for anyx, y∈A.

The next theorem gives a characterization of positive functionals on unital alge- bras.

Theorem 83. Letω be a linear functional onA.

(1) If ω ≥ 0, then kωk = sup{ω(x) :x≥0,kxk ≤1} = limω(hλ) for an ap- proximate unit(hλ)λ∈Λ. In particular, if Ais unitalkωk=ω(1) =|ω(1)|. (2) Ifω1, ω2≥0, thenkω12k=kω1k+kω2k.

(3) IfAis unital, thenkωk=ω(1)⇐⇒ω≥0. Proof. 1) We have

kωk ≥sup{ω(x) :x≥0,kxk ≤1} ≥lim supω(hλ).

Now on the other hand, for everyε >0 there is akyk= 1 such that,

kωk2−ε≤ |ω(y)|2= lim|ω(h1/2λ y)|2≤lim infω(hλ)ω(yy)≤ kωklim infω(hλ).

2) From (1) it follows, that

12k= lim(ω12)(hλ) = limω1(hλ) + limω2(hλ) =kω1k+kω2k.

(20)

3) It suces to show, that ω is positive on the C*-sub-algebras C(x) where x∈A+ ifω(1) =kωk. We can use continuous functional calculus for normal ele- ments, to identifyC(x)∼=C(σ(x)). Therefore we may use the Riesz representation theorem, to identifyωwith a complex measureµ. ω(1) =kωkyieldskµk=µ(σ(x)),

which is only possible forµ≥0. Hence,ω≥0.

An important consequence is, that any state on Auniquely extends to a state on A1.

Fact 84. Let ω be a positive linear functional on the non unital C*-algebra A. Then by ω(1) =kωk, it extends to a positive functional on A1. In particular, the extension of ω∈S(A)belongs toS(A1).

The following estimate, will be useful later on.

Corollary 85. If x∈A, then|ω(x)|2≤ω(xx) for a stateω.

Proof. ω extends toA1. So one can use Lemma (82) withy= 1. 4.3. The GNS Construction. In this section we use the fact, that postive func- tionals induce pre-inner products to obtain representations. For every positive func- tionalω, the so called Gelfand-Naimark-Segal construction, yields a representation πωof the C*-algebra.

Theorem 86. Letω be a positive functional on A.

(1) Nω := {x∈A: ω(xx) = 0} is a left ideal in A. The pre-inner product (., .)ωinduces a well-dened inner product onA/Nωby(x+Nω, y+Nω)ω:=

(x, y)ω. Its completion (Hω,(., .)ω)is a Hilbert space.

(2) Let πω(x)(y+Nω) := xy+Nω for x, y ∈ A. Then πω(x) is a bounded operator on A/Nω with kπω(x)k ≤ kxk. It follows that it extends to a bounded operator onHω which we will also callπω(x).

(3) The mapπω:A→B(Hω)is a representation, called the GNS representa- tion ofAassociated toω.

Proof. 1) Let us start by showing thatNωis a closed left ideal inA. It is obviously closed since it is the zero-set of a continuous map. Fory, z ∈Aand x∈Nω using Lemma (82) we obtain

|ω((yx)yx))|2=|ω(xyyx)|2≤ω(xyy(xyy))ω(xx) = 0.

From|ω(xz)|2≤ω(zz)ω(xx) = 0we infer

ω((x+z)(x+z)) =ω(xx+xz+zx+zz) =ω(xz) +ω(zx) +ω(zz) =ω(zz).

It follows that Nω is a left ideal (for Nω+Nω =Nω one choosesz ∈Nω in the above equation) and by employing the polar formula that the pre-inner product is well dened by acting on the representatives. From

0 =ω((x+Nω)(x+Nω)) =ω((x+Nω)(x+Nω)) =ω(xx)⇒x+Nω= 0 +Nω one concludes that(., .)ωis denit on A/Nω.

2) SinceNω is a left idealπω(x)is a well dened operator onN/Nω. Applying ω to

yxxy≤ kxk2yy yields

ω(x)k ≤ kxk.

(21)

3)πωis obviously a homomorphism. Furthermore

ω(x)(y+Nω), z+Nω)ω=ω(zxy) = (y+Nω, πω(x)(z+Nω))ω

= (πω(x)(y+Nω), z+Nω)ω.

Using Lemma (7), we getπω(x) =πω(x).

We are going to see that representations constructed in this way are always cyclic and unique in a certain sense.

Theorem 87. Let(Hω, πω)be the GNS-representation of Aassociated to ω. (1) There is a vector ξω ∈ Hω satisfying πω(x)ξω = x+Nω and ω(x) =

ω(x)ξω, ξω). ξω is cyclic forπω withkξωk2=kωk.

(2) Let π be a representation ofAonB(H)with aξ cyclic for πand ω(x) = (π(x)ξ, ξ). Then(H, π, ξ)is unitarily equivalent to(Hω, πω, ξω).

Proof. 1) In the unital caseξωis simply1 +Nω. In the general case one considers the extensionω1 ofω toA1 withkω1k=kωk. By Theorem (83),

ω1(1) =kωk= limω(hλ) = limω(h2λ).

Hence,

limk1−hλk2Hω

1 = limω1((1−hλ)2) = 0.

It follows thatA/Nω is dense inA1/Nω1 andHω can be identied withHω1. Now ξωω1 satisesω(x) = (πω(x)ξω, ξω)H.

2) If there was a unitary operator establishing a unitary equivalence, it must certainly sendπ(x)ξtoπω(x)ξω. So we deneU :π(A)H→πω(A)HωbyU π(x)ξ= πω(x)ξω. One calculates

(U π(x)ξ, U π(y)ξ)Hω = (πω(x)ξω, πω(y)ξω)Hω= (πω(yx)ξω, ξω)Hω=

ω(yx) = (π(yx)ξ, ξ)H = (π(x)ξ, π(y)ξ)H.

ThusU is an isometry. In particular, it is well dened. Sinceξ, ξω are cyclic, it uniquely extends to a unitary operator also denoted byU from H to Hω. For all x, y∈A we have

πω(x)U π(y)ξ=πω(xy)ξω=U π(xy)ξ=U π(x)π(y)ξ.

Hence, by a density argumentπω(x)U =U π(x)for allx∈A andU establishes

a unitary equivalence.

Remark 88. As mentioned in Remarks (53, 77) a lot of the things we have done do not work for Banach *-algebras. However, with some eort the results of this section may be obtained for Banach *-algebras with a bounded approximate iden- tity. Central in this considerations is the Cohen-Hewitt Factorisation Theorem. For more details see [4] .

(22)

4.4. The Space of States, Pure States. As mentioned above, so far we did not need all the structure that comes with a C*-algebra. We have seen that for every state there exists a representation but this representation is not necessarily faithful. In order to get a faithful representation we will consider direct sums of GNS representations. If there are suciently many states this approach will work.

Indeed, this is the point where we need most of the structure of C*-algebras.

We will start with a Banach-Alaoglou like theorem.

Theorem 89. IfAis a unital C*-algebra, then (S(A), σ(S(A), ι(A))is a compact convex subset of the dual spaceA.

Proof. The proof works similar to Theorem (25) or Banach-Alaoglou respectively.

Forx∈A+ consider the maps

ιx:A0 →C; φ7→φ(x).

Using these maps, we get

S(A) =K1A(0)∩ {φ∈A:φ(1) = 1} ∩ \

x∈A+

ι−1x ([0,∞)).

Because the ιx are continuous, this is an intersection of w*-closed sets. Hence, S(A) is also w*-closed. By Banach-Alaoglou, K1A(0)is compact. Hence S(A) is compact.

To see, thatS(A)is convex, note, that the sum of positive functionals is always a positive functional. Moreover, by Theorem (83)

kλω1+ (1−λ)ω2k=λkω1k+ (1−λ)kω2k= 1

ifω1, ω2∈S(A). Therefore, a convex combination of states is always a state.

The next theorem one could understand as a version of the algebraic Hahn- Banach theorem for states.

Theorem 90. If B ⊆A is C*-sub-algebra and ω is a state on B, then ω extends to a state ω0 on A.

Proof. By the Hahn-Banach theorem there is an extension ω0 with norm 1. If B is unital, then by Theorem (83)ω0 is a state since ω0(1) =ω(1) = 1. If B is not unital, one extends ω to B1 ⊆ A1 and then there is an extension ω0 on A1. By

restricting toAone gets the wanted extension.

If the set of positive functionals would fully describe a C*-algebra, obviously the subset of states also would. As it will eventually turn out, there is an even smaller subset sucient for that task.

Denition 91. A state ω is called pure if it is an extrem point of S(A), i.e.

ω=λω1+ (1−λ)ω2with 0< λ <1 andω1, ω2∈S(A), yields ω12=ω. The set of pure states is denotedP(A).

Lemma 92. A state ω is pure, if and only if all positive ω with ω ≤ ω are multiples ofω. Hereω≤ω mean, that ω(x)≤ω(x)for allx∈A+.

Proof. Letω be pure andω ≤ω. Since kω−ωk=kωk − kωk (cf. Theorem (83),(2)), we may assume, that1 =kωk>kωk. We get

ω=kωk ω

k + (1− kωk) (ω−ω) (1− kωk)

Referenzen

ÄHNLICHE DOKUMENTE

3 The theory of Dieudonn´ e modules is still today an active field of research, together with formal groups and p-divisible groups (work of Fontaine, Messing, Zink.. basis, and

A cofibrantly generated model category is called weakly finitely generated if the domains and the codomains of I are finitely presentable, the domains of the maps in J are small, and

Gruppenalgebran Gber nicht--zyklischen p-Gruppen~ J.Reine Ang.Math.. The indecozposable representations of the dihedral

In this section we are going to recall the semi-model structures on the categories of operads and operadic algebras in a monoidal model category.. After that we will state a version

Our main theorem says that if the condition for the right von Neumann algebra to be the commutant of the left one holds, both von Neumann algebras are type III 1 factors, according

Lemma: The functor Fj induces an equivalence between Pj and R(Sj).. Moreover, in that case each inde- composable has a peak. To apply these results, we have to

There are only finitely many multiplicity-free tilting modules T A with End(T A) of finite representation type, and any such tilting module has both

Dually, in the case when a splitting filtration is given by Proposition 1* (such as the dual Rojter filtration, or any filtration derived from the preprojective partition),