(N.P. Landsman) Lecture Notes On Operator Algebra
(N.P. Landsman) Lecture Notes On Operator Algebra
Operator Algebras
N.P. Landsman
Institute for Mathematics, Astrophysics, and Particle Physics
Radboud University Nijmegen
Heyendaalseweg 135
6525 AJ NIJMEGEN
THE NETHERLANDS
email: [email protected]
website: http://www.math.ru.nl/∼landsman/OA2011.html
tel.: 024-3652874
office: HG03.740
1 HISTORICAL NOTES AND OVERVIEW 2
• Hilbert spaces (John von Neumann),1 leading to the theory of von Neumann
algebras (originally called rings of operators by von Neumann himself).
So, roughly speaking, the theory of operator algebras is the same as the theory of
von Neumann algebras and C ∗ -algebras.3 Let us elaborate on each of these in turn.
concept of a matrix; in previous research he had even used infinite matrices.4 Born
turned to his former teacher Hilbert for mathematical advice. Hilbert had been
interested in the mathematical structure of physical theories for a long time; his
Sixth Problem (1900) called for the mathematical axiomatization of physics. Aided
by his assistants Nordheim and von Neumann, Hilbert thus ran a seminar on the
mathematical structure of quantum mechanics, and the three wrote a joint paper
on the subject (which is now obsolete).
It was von Neumann alone who, at the age of 23, recognized the mathematical
structure of quantum mechanics. In this process, he defined the abstract concept
of a Hilbert space, which previously had only appeared in some examples. These
examples went back to the work of Hilbert and his pupils in Göttingen on integral
equations, spectral theory, and infinite-dimensional quadratic forms. Hilbert’s fa-
mous memoirs on integral equations had appeared between 1904 and 1906. In 1908,
his student E. Schmidt had defined the space `2 in the modern sense, and F. Riesz
had studied the space of all continuous linear maps on `2 in 1912. Various examples
of L2 -spaces had emerged around the same time. However, the abstract notion of a
Hilbert space was missing until von Neumann provided it.
Von Neumann saw that Schrödinger’s wave functions were unit vectors in a
Hilbert space of L2 type, and that Heisenberg’s observables were linear operators
on a different Hilbert space, of `2 type. A unitary transformation between these
spaces provided the the mathematical equivalence between wave mechanics and
matrix mechanics. (Similar, mathematically incomplete insights had been reached
by Pauli and Dirac.) In a series of papers that appeared between 1927–1929, von
Neumann defined Hilbert space, formulated quantum mechanics in this language,
and developed the spectral theory of bounded as well as unbounded normal operators
on a Hilbert space. This work culminated in his book Mathematische Grundlagen
der Quantenmechanik (1932), which to this day remains the definitive account of
the mathematical structure of elementary quantum mechanics.5
More precisely, von Neumann proposed the following mathematical formulation
of quantum mechanics (cf. Ch. 2 below for minimal background on Hilbert spaces).
1. The observables of a given physical system are the self-adjoint linear opera-
tors a on a Hilbert space H.
2. The states of the system are the so-called density operators ρ̂ on H, that
is, the positive trace-class operators on H with unit trace.
As a special case, take a unit vector Ψ in H and form the associated projection
pΨ = |ΨihΨ|, (1.2)
where we use the following notation (due to Dirac): for any two vectors Ψ, Φ in H,
the operator |ΨihΦ| is defined by6
|ΨihΦ|Ω = (Φ, Ω)Ψ. (1.3)
In particular, if Φ = Ψ is a unit vector, then (1.2) is the projection onto the one-
dimensional subspace C · Ψ of H spanned by Ψ (see Exercise 1 below for a rappèl
on projections). In that case, it is easily shown (see Exercise 2) that the density
operator ρ̂ = pΨ leads to
< a >pΨ ≡< a >Ψ = Tr (pΨ a) = (Ψ, aΨ). (1.4)
Special states like pΨ (often confused with Ψ itself, which contains additional phase
information) are called pure states, whereas all other states are said to be mixed.
Let B(H) be the space of all bounded operators on H, with unit operator simply
denoted by 1. A functional ω : B(H) → C (which is linear by definition) is called:
• positive when ω(a∗ a) ≥ 0 for all a ∈ B(H);
• normalized when ω(1) = 1.
We will see that positivity implies continuity (see Exercise 3). For reasons to become
clear shortly, a normalized positive functional on B(H) is called a state on B(H);
our earlier use of this word should accordingly be revised a little. Indeed, it is trivial
to show that for any density operator ρ̂, the functional ρ : B(H) → C defined by
ρ(a) = Tr (ρ̂a) (1.5)
is a state on B(H). Conversely, are all states on B(H) of this kind?
Von Neumann implicitly assumed a certain continuity condition on states, which
in modern terminology is called σ-weak (or ultraweak) continuity, which implies
that the answer is yes; states à la (1.5) (in other words, σ-weakly continuous states)
are called normal states on B(H). However, without this continuity condition the
set of states on B(H) turns out not to be exhausted by density operators on H
(unless H is finite-dimensional), although it is hard to give explicit examples.
The set of all states on B(H) is obviously convex (within B(H)∗ ), as is its subset
of all normal states on B(H). The extreme boundary of a convex set K is the set
of all ω ∈ K that are indecomposable, in the sense that if ω = λω1 + (1 − λ)ω2 for
some 0 < λ < 1 and ω1 , ω2 ∈ K, then ω1 = ω2 = ω.7 Von Neumann saw that states
(1.4) precisely correspond to the points in the extreme boundary of the convex set
Sn (B(H)) of normal states on B(H): on other words, a density operator ρ̂ yields an
extreme point ρ in Sn (B(H)) iff it is of the form (1.4). See also Exercise 4.
6
Dirac wrote |Ωi for Ω, etc., which is superfluous, but in this special case it leads to the neater
expression |ΨihΦ|Ωi = hΦ|Ωi|Ψi, where the inner product is written as hΦ|Ωi ≡ (Φ, Ω).
7
In some examples of compact convex sets in Rn , the extreme boundary of K coincides with
its geometric boundary; cf. the closed unit ball. However, the extreme boundary of an equilateral
triangle consists only of its corners. If K fails to be compact, its extreme boundary may even be
empty, as illustrated by the open unit ball in any dimension.
1 HISTORICAL NOTES AND OVERVIEW 5
1.3 C*-algebras
Following the pioneering work of von Neumann, an important second step in the the-
ory of operator algebras was the initiation of the theory of C ∗ -algebras by Gelfand
and Naimark in 1943. It turns out that von Neumann’s “rings of operators” are
special cases of C ∗ -algebras, but von Neumann algebras also continue to be studied
on their own. A fruitful mathematical analogy is that C ∗ -algebras provide a non-
commutative generalization of topology, whereas von Neumann algebras comprise
noncommutative measure theory.9 To understand this, we state the most important
fact about C ∗ -algebras, namely that there are two totally different way of approach-
ing them:
We start with the second. In 1943, Gelfand and Naimark noted that the space C(X)
has the following additional structure beyond just being a commutative algebra over
C (see Exercises). Firstly, it has a norm, given by (a ∈ C(X))
where we have written k·k for k·k∞ , and a, b ∈ C(X). But more structure is needed!
An involution on an algebra A is a real-linear map A → A∗ such that a∗∗ = a,
(ab)∗ = b∗ a∗ , and (λa)∗ = λa∗ for all a, b ∈ A and λ ∈ C. An algebra with involution
is also called a ∗ -algebra. Secondly, then, C(X) has an involution a 7→ a∗ , given
by a∗ (x) = a(x). This involution is related to the norm as well as to the algebraic
structure by the property
ka∗ ak = kak2 . (1.8)
We summarize these properties by saying that C(X) is a commutative C ∗ -algebra
with unit, in the following sense:
2. Prove (1.4).
3. Prove that a state ω on a C ∗ -algebra A (with unit) satisfies |ω(a)| ≤ kak (and
hence is continuous with norm kωk = ω(1)).
x2 + y 2 + z 2 ≤ 1.
(c) Conclude that the state space S(M2 (C)) of M2 (C) is isomorphic (as a
convex set) to the three-ball B 3 = {(x, y, z) ∈ R3 | x2 + y 2 + z 2 ≤ 1}.
(d) Under this isomorphism, show that the extreme boundary of S(M2 (C))
corresponds to the two-sphere S 2 = {(x, y, z) ∈ R3 | x2 + y 2 + z 2 = 1}.
(e) Verify that the states in the extreme boundary of S(M2 (C)) are exactly
those of the form (1.4), where Ψ ∈ C2 is a unit vector.
2.2 Completeness
Many concepts of importance for Hilbert spaces are associated with the metric rather
than with the underlying inner product or norm. The main example is convergence:
Definition 2.3 1. Let (xn ) := {xn }n∈N be a sequence in a metric space (V, d).
We say that xn → x for some x ∈ V when limn→∞ d(xn , x) = 0, or, more
precisely: for any ε > 0 there is N ∈ N such that d(xn , x) < ε for all n > N .
In a normed space, hence in particular in a space with inner product, this
therefore means that xn → x if limn→∞ kxn − xk = 0.
Clearly, a convergent sequence is Cauchy: from the triangle inequality and symmetry
one has d(xn , xm ) ≤ d(xn , x) + d(xm , x), so for given ε > 0 there is N ∈ N such that
d(xn , x) < ε/2, et cetera. However, the converse statement does not hold in general,
as is clear from the example of the metric space (0, 1) with metric d(x, y) = |x − y|:
the sequence xn = 1/n does not converge in (0, 1). In this case one can simply
extend the given space to [0, 1], in which every Cauchy sequence does converge.
Definition 2.4 A metric space (V, d) is called complete when every Cauchy se-
quence in V converges (i.e., to an element of V ).
• A vector space with norm that is complete in the associated metric is called a
Banach space. In other words: a vector space B with norm k · k is a Banach
space when every sequence (xn ) such that limn,m→∞ kxn − xm k = 0 has a limit
x ∈ B in the sense that limn→∞ kxn − xk = 0.
• A vector space with inner product that is complete in the associated metric is
called a Hilbert space. In other words: a vector space H with innerpproduct
( , ) is a Hilbert space when it is a Banach space in the norm kxk = (x, x).
K ⊥ := {f ∈ H | f ⊥ K}. (2.15)
2. One has
K ⊥⊥ := (K ⊥ )⊥ = K. (2.17)
Theorem 2.7 Two Hilbert spaces are isomorphic iff they have the same dimension.
15
By definition of the norm, if f ⊥ g one has Pythagoras’ theorem kf + gk2 = kf k2 + kgk2 .
16
A subspace of a vector space is by definition a linear subspace.
2 REVIEW OF HILBERT SPACES 12
The functions (δs )s∈S , defined by δs (t) = δst , t ∈ S, clearly form an o.n.b. of `2 (S).
Now let H be an n-dimensional Hilbert space; a case in point is H = Cn .
By definition, H has an o.n.b. (ei )ni=1 . Take S = n = {1, 2, . . . , n}. The map
u : H → `2 (n), given by linear extension of uei = δi is unitary and provides an
isomorphism H ∼ = `2 (n). Hence all n-dimensional Hilbert space are isomorphic.
• If S is countable, then `2 (S) = {f : S → C | kf k2 < ∞}, with
!1/2
X
kf k2 := |f (s)|2 , (2.19)
s∈S
with inner product given by (2.18); this is finite for f, g ∈ `2 (S) by the Cauchy–
Schwarz inequality. Once again, the functions (δs )s∈S form an o.n.b. of `2 (S), and
the same argument shows that all separable Hilbert space are isomorphic to `2 (N)
and hence to each other. A typical example is `2 (Z).
• If S is uncountable, then `2 (S) is defined as in the countable case, where
the sum in (2.19) is now defined as the supremum of the same expression
evaluated on each finite subset of S. Similarly, the sum in (2.18) is defined by
first decomposing f = f1 − f2 + i(f3 − f4 ) with fi ≥ 0, and g likewise; this
decomposes (f, g) as a linear combination of 16 non-negative terms (fi , gj ),
each of which is defined as the supremum over finite subsets of S, as for kf k2 .
The previous construction of an o.n.b. of `2 (S) still applies verbatim, as does the
proof that any Hilbert space of given cardinality is isomorphic to `2 (S) for some S
of the same cardinality. In sum, we have proved (von Neumann’s) Theorem 2.7.
Let us note that for infinite sets S we may regard `2 (S) as the closure in the
norm (2.19) of the (incomplete) space `c (S) of functions that are nonzero at finitely
many s ∈ S; this means that for any f ∈ `2 (S) there is a sequence (fn ) in `c (S) such
that limn→∞ kfn − f k2 = 0. In what follows, we also encounter the Banach space
`∞ (S) = {f : S → C | kf k∞ < ∞}; (2.20)
kf k∞ := sup{|f (s)|}, (2.21)
s∈S
For any p > 0, we define Lp (X, µ) as the space of Borel functions20 on X for which
Z 1/p
p
kf kp := dµ |f | < ∞, (2.30)
X
where similarly ambiguous notation has been used as for L2 (Rn ) (cf. the end of §2.5).
20
Here f : X → C is Borel when fi−1 ((s, t)) ∈ B(X) for each 0 ≤ s < t, i = 1, 2, 3, 4, where
f = f1 − f2 + i(f3 − f4 ) is the unique decomposition with fi ≥ 0 (e.g., f1 (x) = max{Re(f (x)), 0}).
2 REVIEW OF HILBERT SPACES 15
Note that a 7→ a∗ is anti-linear: one has (λa)∗ = λa for λ ∈ C. Also, one has
The adjoint allows one to define the following basic classes of bounded operators:
1. n : H → H is normal when n∗ n = nn∗ .
2. a : H → H is self-adjoint when a∗ = a (hence a is normal).
3. a : H → H is positive, written a ≥ 0, when (f, af ) ≥ 0 for all f ∈ H.
4. p : H → H is a projection when p2 = p∗ = p (hence p is positive).
5. u : H → H is unitary when u∗ u = uu∗ = 1 (hence u is normal).
6. v : H → H is an isometry when v ∗ v = 1, and a partial isometry when v ∗ v
is a projection (in which case vv ∗ is automatically a projection, too).
The following notion plays a fundamental role in von Neumann algebra theory:
Definition 2.20 A trace-class operator ρ : H → H is called a density
P operator if
ρ is positive and Tr (ρ) = 1 (so that kρk1 = 1).
P Equivalently, ρ = i λi pi (strongly)
with dim(pi ) < ∞ for all i, 0 < λi ≤ 1, and i λi = 1.
3 C*-ALGEBRAS 19
3 C*-algebras
3.1 Basic definitions
If a and b are bounded operators on H, then so is their sum a + b, defined by
(a + b)(v) = av + bv, and their product ab, given by (ab)(v) = a(b(v)). This
follows from the triangle inequality for the norm and from (2.34), respectively. Also,
homogeneity of the norm yields that ta is bounded for any t ∈ C. Consequently, the
set B(H) of all bounded operators on a Hilbert space H forms an algebra over the
complex numbers, having remarkable properties. To begin with (cf. (2.32)):
Proposition 3.1 The space B(H) of all bounded operators on a Hilbert space H is
a Banach space in the operator norm
kak := sup {kaf kH , f ∈ H, kf kH = 1}. (3.1)
This is a basic result from functional analysis; it even holds if H is a Banach space.
Definition 3.2 A Banach algebra is a Banach space A that is simultaneously an
algebra in which kabk ≤ kak kbk for all a, b ∈ A.
According to (2.34), we see that B(H) is not just a Banach space but even a Banach
algebra. Also this would still be the case if H were merely a Banach space, but the
fact that it is a Hilbert space gives a crucial further ingredient of the algebra B(H).
Definition 3.3 1. An involution on an algebra A is a real-linear map A → A∗
such that a∗∗ = a, (ab)∗ = b∗ a∗ , and (λa)∗ = λa∗ for all a, b ∈ A and λ ∈ C.
An algebra with involution is also called a ∗ -algebra.
2. A C ∗ -algebra is a Banach algebra A with involution in which for all a ∈ A,
ka∗ ak = kak2 . (3.2)
3.3 States
The concept of a state originates with quantum physics, but also purely mathemat-
ically it came to play a dominant (and beautiful) role in operator algebra theory.
Definition 3.5 A state on a unital C ∗ -algebra A is a linear map ω : A → C that
is positive, in that ω(a∗ a) ≥ 0 for all a ∈ A, and normalized, in that ω(1) = 1.
If we define the dual A∗ of A as the space of linear maps ϕ : A → C for which
kϕk = sup{|ϕ(a)|, a ∈ A, kak = 1} (3.6)
is finite (cf. (2.32)), then it can be shown that any state ω on A lies in A∗ , with
kωk = 1. This leads to an extension of Definition 3.5 to general (i.e., not necessarily
unital) C ∗ -algebras: a state on a C ∗ -algebra A is a functional ω : A → C that is
positive and normalized in the sense that kωk = 1. This implies ω(1) = 1 whenever
A does have a unit, so that the two definitions are consistent when they overlap.
The state space S(A) of A (i.e., the set of all states on A) is a convex set: if
ω1 and ω2 are states, then so is λω1 + (1 − λ)ω2 for any λ ∈ [0, 1]. It follows that
if
P(ω1 , ω2 , . . . , ωn P
) are states, and (λ1 , λ2 , . . . , λn ) are numbers in [0, 1] such that
λ
i i = 1, then i λi ωi is a state. This extends to infinite sums if we equip S(A)
with the weak topology inherited from A∗ (in which ωn → ω if ωn (a) → ω(a) for
∗
each a ∈ A). If A has a unit, then S(A) is a compact convex set in this topology.29
Definition 3.6 A state ω is pure if ω = λ ω1 + (1 − λ)ω2 for some λ ∈ (0, 1) and
certain states ω1 and ω2 implies ω1 = ω2 . The pure states on A comprise the pure
state space of A, denoted by P (A) or ∂S(A). If a state is not pure, it is mixed.30
The convex structure of the state space is nicely displayed in the noncommutative
case by A = M2 (C), the C ∗ -algebra of 2 × 2 complex matrices. Put
1 + z x + iy
ρ = 12 ; (3.7)
x − iy 1 − z
then ρ is a density matrix on C2 iff (x, y, z) ∈ R3 with x2 + y 2 + z 2 ≤ 1; this set is
the three-ball B 3 in R3 . It is easy to see that ρ defines a state ωρ on the M2 (C) by
ωρ (a) = Tr (ρa). (3.8)
Conversely, every state on M2 (C) is of this form, so that the state space S(M2 (C))
is isomorphic (as a convex set) to B 3 . The pure states ∂B 3 then correspond to the
two-sphere S 2 = {(x, y, z) ∈ R2 | x2 + y 2 + z 2 = 1} (see exercises).
In the commutative case, we have the key behind the proof of Theorem 3.4:
Lemma 3.7 The pure state space P (A) ⊂ S(A) ⊂ A∗ of a commutative C ∗ -algebra
A coincides with its Gelfand spectrum Σ(A) (seen as a subspace of A∗ ).
The proof is an exercise, but one can see the point from the example A = C(X).
29
This follows from the Banach–Alaoglu Theorem of functional analysis; see [8] or exercises.
30
The Krein–Milman Theorem of functional analysis [8] guarantees the abundance of pure states
in compact convex sets in that any state is a convex sumof pure states (or limit thereof).
3 C*-ALGEBRAS 22
3.4 Spectrum
To prove isometry of the Gelfand transform (and hence Theorem 3.4) from Lemma
3.7, we need a nice result with an ugly proof based on the Axiom of Choice (ac):
Lemma 3.8 Let A be a C ∗ -algebra with unit. For any self-adjoint a ∈ A, there is
a pure state ω0 ∈ P (A) such that |ω0 (a)| = kak.
This will be proved in a minute; for now, we just point out that for A = C(X),
X compact, this is immediate from Weierstrass’ Theorem stating that a continuous
function on a compact set assumes its maximum (and its minimum). Given Lemma
3.8, if a∗ = a, then kak = |ω0 (a)| = |â(ω0 )| ≤ kâk∞ ≤ kak, the last inequality
arising because kâk∞ = sup{|â(ω)|, ω ∈ Σ(A)}, (3.5), and |ω(a)| ≤ kak (since ω is
a state). Hence kâk∞ = kak for self-adjoint a, and therefore for any a (exercise).
The proof of Theorem 3.4 is now complete up to Lemma 3.8. To prove the latter,
and for many other reasons, we introduce the following extremely important notion.
Definition 3.9 Let A be a Banach algebra with unit. The spectrum σ(a) of a ∈ A
is the set of all z ∈ C for which a − z ≡ a − z · 1 has no (two-sided) inverse in A.
The spectral radius r(a) of a ∈ A is defined as r(a) := sup{|z|, z ∈ σ(a)}.
We quote two basic results from functional analysis [8]:31
Proposition 3.10 1. The spectrum σ(a) is a nonempty compact subset of C.
Σ(C(X)) ∼
= X, (3.9)
Lemma 3.15 Let A = C0 (X) for some noncompact locally compact Hausdorff space
X. Then Ȧ ∼ = C(Ẋ), where 1 ∈ Ȧ is identified with the constant function 1Ẋ
in C(Ẋ). Conversely, removing C1Ẋ from C(Ẋ) corresponds to removing C from
Ȧ = A ⊕ C (as a vector space), leaving one with C0 (X).
ω∞ (a + λ1) := λ, (3.18)
is a character of Ȧ.
We may now prove Theorem 3.4 also in the nonunital case. Applying the unital
case of Theorem 3.4 to Ȧ and using Lemma 3.16, one finds Ȧ ∼ = C(Ẋ) with X :=
Σ(A). Formally, we now use a little lemma stating that if A and B are C ∗ -algebras
without unit, then Ȧ ∼= Ḃ iff A ∼
= B. Informally: removing C from Ȧ = A ⊕ C
precisely leaves one with C0 (X) by Lemma 3.15, so that finally A ∼
= C0 (X).
Using the invariance of Haar measure, it is trivial to verify that this product is com-
mutative if G is abelian. We also define an involution on Cc (G) by f ∗ (x) = f (−x),
where −x ≡ x−1 .39 Of course, we would now like to turn Cc (G) into a commutative
C ∗ -algebra, but the most obvious norms like the Lp -ones do not accomplish this.
Instead, for f ∈ Cc (G) we define an operator π(f ) on the Hilbert space L2 (G)
by π(f )ψ = f ∗ ψ; here we initially pick ψ ∈ Cc (G) and show that
Z
kπ(f )k ≤ kf k1 := dx |f (x)|, (3.20)
G
so that π(f ) is bounded and may be extended to all of L2 (G) by continuity. As-
sociativity of convolution then implies π(f ∗ g) = π(f )π(g), and also one has
π(f ∗ ) = π(f )∗ . The map f 7→ π(f ) from Cc (G) to B(L2 (G)) is injective (exer-
cise), so that kf k = kπ(f )k defines a norm on Cc (G). One immediately sees that
the axioms (1.7) and (1.8) are satisfied, so that the completion of Cc (G) in this
norm, called C ∗ (G), is a commutative C ∗ -algebra.40 What is its Gelfand spectrum?
Recall that, for any locally compact abelian group G, the dual group or character
group Ĝ is defined as Ĝ = Hom(G, T), i.e., the continuous group homomorphisms
from G to T, equipped with the compact-open topology.41 For example, for G = R
we have Ĝ ∼= R, where p ∈ R defines a character χp ∈ R̂ by χp (x) = exp(ipx). On
the other hand, for G = T one finds Ĝ ∼= Z, where n ∈ Z defines χn (z) = z n , z ∈ T.
Theorem 3.17 Let G be a locally compact abelian group. The Gelfand spectrum
Σ(C ∗ (G)) is homeomorphic to Ĝ, so that C ∗ (G) ∼
= C0 (Ĝ), and the Gelfand transform
ˆ
f 7→ f implementing this isomorphism coincides with the Fourier transform
Z
fˆ(χ) = dx χ(x)f (x). (3.21)
G
4. Prove:
(a) Σ(C ∗ (G)) = Σ(L1 (G)), where the Banach algebra L1 (G) is the comple-
tion of Cc (G) in the L1 -norm k · k1 .
(b) Σ(L1 (G)) ∼ = Ĝ, first by proving that each ωχ is a character of L1 (G),
χ ∈ Ĝ, and secondly that there are no others.
3 C*-ALGEBRAS 29
(a) Show that the inner product (pω a, pω b) = ω(a∗ b) on A/Nω is well defined;
(b) Show that the representation πω (a)pω b := pω ab is well defined on A/Nω ;
(c) Show that πω (a) is bounded.
We first show that the sum is a Cauchy sequence. Indeed, for n > m one has
n
X m
X n
X n
X n
X
k k k k
k a − a k=k a k ≤ ka k ≤ kakk .
k=0 k=0 k=m+1 k=m+1 k=m+1
Hence n
X
k1 − ak (1 − a)k = kan+1 k ≤ kakn+1 .
k=0
Pn
which → 0 for n → ∞, as kak < 1 by assumption. Thus limn→∞ k=0 ak (1−a) = 1.
By a similar argument,
n
X
lim (1 − a) ak = 1.
n→∞
k=0
Hence a + b = a(1 + a−1 b) has an inverse, namely (1 + a−1 b)−1 a−1 , which exists by
(3.27) and Lemma 3.27. It follows that all c ∈ A for which ka − ck < lie in G(A),
for ≤ ka−1 k−1 .
3 C*-ALGEBRAS 35
Hence ∞
X
g(z) = (z0 − z)k (z0 − a)−k−1 (3.28)
k=0
Now suppose that σ(a) = ∅, so that ρ(a) = C. The function g, and hence gρ , is
then defined on C, where it is analytic and vanishes at infinity. In particular, gρ is
bounded, so that by Liouville’s theorem it must be constant. By (3.30) this constant
is zero, so that g = 0.44 This is absurd, so that ρ(a) 6= C hence σ(a) 6= ∅. This
finishes the proof of Proposition 3.10.1.
44
This follows by a basic result in Banach spaces B: if v ∈ B is such that ρ(v) = 0 for all ρ ∈ B ∗ ,
then v = 0.
3 C*-ALGEBRAS 36
The proof of Proposition 3.10.2 is as follows. By Lemma 3.27, for |z| > kak the
function g in the proof of Lemma 3.28 has the norm-convergent power series
∞
1 X a k
g(z) = . (3.31)
z k=0 z
On the other hand, we have seen that for any z ∈ ρ(a) one may find a z0 ∈ ρ(a)
such that the power series (3.28) converges. If |z| > r(a) then z ∈ ρ(a), so (3.28)
converges for |z| > r(a). At this point the proof relies on the theory of analytic
functions with values in a Banach space, which says that, accordingly, (3.31) is norm-
convergent for |z| > r(a), uniformly in z. Comparing with (3.26), this sharpens what
we know from Lemma 3.27. The same theory says that (3.31) cannot norm-converge
uniformly in z unless kan k/|z|n < 1 for large enough n. This is true for all z for
which |z| > r(a), so that
lim sup kan k1/n ≤ r(a). (3.32)
n→∞
To derive a second inequality we use the following polynomial spectral mapping
property.
Lemma 3.29 For a polynomial p on C, define p(σ(a)) as {p(z)| z ∈ σ(a)}. Then
p(σ(a)) = σ(p(a)). (3.33)
To prove this equality, choose z, α ∈ C and compare the factorizations
n
Y
p(z) − α = c (z − βi (α));
i=1
n
Y
p(a) − α1 = c (a − βi (α)1). (3.34)
i=1
Here the coefficients c and βi (α) are determined by p and α. When α ∈ ρ(p(a))
then p(a) − α1 is invertible, which implies that all a − βi (α)1 must be invertible.
Hence α ∈ σ(p(a)) implies that at least one of the a − βi (α)1 is not invertible, so
that βi (α) ∈ σ(a) for at least one i. Hence α ∈ p(βi (α)) − α = 0, i.e., α ∈ p(σ(a)).
This proves the inclusion σ(p(a)) ⊆ p(σ(a)).
Conversely, when α ∈ p(σ(a)) then α = p(z) for some z ∈ σ(a), so that for some
i one must have βi (α) = z for this particular z. Hence βi (α) ∈ σ(a), so that a−βi (α)
is not invertible, implying that p(a) − α1 is not invertible, so that α ∈ σ(p(a)). This
shows that p(σ(a)) ⊆ σ(p(a)), and (3.33) follows.
To conclude the proof of Proposition 3.10.2, we note that since σ(a) is closed
there is an α ∈ σ(a) for which |α| = r(a). Since αn ∈ σ(an ) by Lemma 3.29, one has
|αn | ≤ kan k by (3.26). Hence kan k1/n ≥ |α| = r(a). Combining this with (3.32)
yields
lim sup kan k1/n ≤ r(a) ≤ kan k1/n .
n→∞
Hence the limit must exist, and
lim kan k1/n = inf kan k1/n = r(a).
n→∞ n
4 VON NEUMANN ALGEBRAS 37
Theorem 4.1 Let M be a unital ∗ -subalgebra of B(H). Then the following cond-
tions are equivalent (and hence each defines M to be a von Neumann algebra):
(1) M 00 = M ;
(2) M is closed in the weak operator topology;
(3) M is closed in the strong operator topology;
(4) M is closed in the σ-weak operator topology.
Since the (Lebesgue) integral appearing here is defined for the far larger class B(σ(a))
of bounded Borel functions on σ(a), we may try to extend the continuous functional
calculus f 7→ f (a) to a “Borel functional calculus.” To do so, we regard B(σ(a)) as
a commutative C ∗ -algebra under pointwise operations and the sup-norm, and intro-
duce the commutative von Neumann (and hence C ∗ ) algebra W ∗ (a) = C ∗ (a, 1)00 .
Theorem 4.3 Let a∗ = a ∈ B(H). The isomorphism C(σ(a)) → C ∗ (a, 1) of Theo-
rem 3.18 has a unique extension to a homomorphism B(σ(a)) → W ∗ (a) that satisfies
kf (a)k ≤ kf k∞ for each f ∈ B(σ(a)), and (4.35) remains valid for such f .
The proof [28, Thm. 4.5.4] is based on the following fact [28, Lemma 3.2.2].
Lemma 4.4 There is a bijective correspondence between bounded sesquilinear forms
Q : H × H → C on a Hilbert space H and bounded operators on H, as follows:
• Any b ∈ B(H) defines a bounded sesquilinear form Qb by Qb (v, w) = (v, bw),
which satisfies the bound |Qb (v, w)| ≤ Ckvkkwk (indeed, take C = kbk).
• Conversely, for any sequilinear form Q that satisfies |Q(v, w)| ≤ Ckvkkwk
there is unique b ∈ B(H) such that Q = Qb , with kbk ≤ C.
It suffices to define Q on the diagonal in H × H, that is, Q(v,
Pw) is determined by
3
Q(v, v) ≡ Q(v) through the polarization identity Q(v, w) = 4 k=0 ik Q(w + ik v).
1
This correspondence applies to (4.35) byR taking the special case ρ̂ = |ΨihΨ|, so
that ρ(b) ≡ ψ(b) = (Ψ, bΨ): if Q(Ψ) = σ(a) dµψ f , then Q = Qf (a) , at least for
f ∈ C(σ(a)). The point now is that for f ∈ B(σ(a)) the (Lebesgue) integral on the
right-hand side of (4.35) remains well defined, yielding a sesquilinear form bounded
by C = kf k∞ . This, then, defines the operator f (a) through Q = Qf (a) .
To prove that f (a) ∈ W ∗ (a), we use a much deeper lemma [28, Prop. 6.2.9]:47
Lemma 4.5 B(σ(a), R) is the bounded monotone sequential completion of C(σ(a), R).
Now define B! = {f ∈ B(σ(a), R) | f (a) ∈ W ∗ (a)}, so C(σ(a), R) ⊆ B! ⊆ B(σ(a), R).
If (fn ) is some bounded monotone sequence in B! with pointwise limit f , then—since
the map f 7→ f (a) is a homomorphism and hence preserves positivity—(fn (a)) is a
bounded monotone sequence in B(H), which strongly converges to f (a) [28, Prop.
4.5.2]. Since W ∗ (a) is closed under strong limits, it follows that B! is monotone
sequentially complete, so that by Lemma 4.5, B! = B(σ(a), R). Complexifying, we
obtain f (a) ∈ W ∗ (a) for all f ∈ B(σ(a)). The rest of the proof is an exercise.
47
I.e., B(σ(a), R) is the smallest space of bounded real functions on σ(a) that contains C(σ(a), R)
and is closed under bounded pointwise limits of monotone (increasing or decreasing) sequences.
4 VON NEUMANN ALGEBRAS 39
This is the second claim. The first equality sign is true, because if Y is a closed
subspace of a Banach space Y , then (X/Y )∗ = {ϕ ∈ X ∗ | ϕ Y = 0}.
For the remainder of the theorem, note that aλ → a σ-weakly in M whenever
ρ(aλ − a) → 0 for all ϕ ∈ B(H)∗ . By (4.38), this is equivalent to aλ → a in the
weak∗ -topology, since a possible component of ϕ in M ⊥ drops out.
Corollary 4.11 Each ϕ ∈ M∗ is of the form ϕ(a) = Tr (ρ̂a), for some ρ̂ ∈ B1 (H).
49
A proof of this theorem will be sketched Appendix 3.
4 VON NEUMANN ALGEBRAS 41
2. Derive the following useful reformulation of Theorem 4.1 from the latter:
Let M be a unital ∗ -subalgebra of B(H). Then the closures of M in the strong,
weak, and σ-weak topologies coincide with each other and with M 00 .
4. In the notation of the proof of Theorem 4.1, prove that (Mn )00 = (M 00 )n .
(a) Show that the σ-weak topology on B(H) is the relative weak topology
on B(H)∞ (i.e., the weak topology on B(H ∞ ) restricted to B(H)∞ ).
(b) Define a new topology on B(H), called the σ-strong topology, by re-
stricting the strong topology on B(H ∞ ) to B(H)∞ . Use the same trick
as above (i.e., the passage of H to H d ) to prove the following version of
the Double Commutant Theorem (and hence (1) ⇔ (4) in Theorem 4.1):
Theorem 4.12 Let M be a unital ∗ -algebra in B(H). The following
conditions are equivalent:
i. M is a von Neumann algebra;
ii. M is closed in the σ-weak operator topology;
iii. M is closed in the σ-strong operator topology.
7. Complete the proof of Theorem 4.3 by showing that the map f 7→ f (a),
initially from B(σ(a)) to B(H), is a homomorphism of C ∗ -algebras.
The claim about the norm-topology follows from Proposition 3.21, for von Neumann
algebras are C ∗ -algebras. Since ϕ is isometric, it induces a dual isomorphism (of
Banach spaces) ϕ∗ : N ∗ → M ∗ , with the property that M ∼ = (ϕ∗ (N∗ ))∗ under
∗
the map a 7→ (ϕ (ω) 7→ ω(ϕ(a))), a ∈ M , ω ∈ N∗ . Uniqueness of the predual
(cf. [33, Vol. I Cor. iii.3.9]) then yields ϕ∗ (N∗ ) ∼
= M∗ , which in turn implies that
ϕ preserves pointwise convergent nets: if ω (aλ ) → ω 0 (a) for all ω 0 ∈ M∗ , then
0
A second instructive proof is based on the projection lattice; cf. Theorem 4.8.
Theorem 4.13 then follows from two key properties of completely additive maps:
3. Hence W ∗ (a) ∼
= B(σ(a))/ ker Ba ∼
= L∞ (σ(a), µ) as Banach spaces (via Ba ).
Theorem 4.18 Let M ⊂ B(H) be an abelian von Neumann algebra (H separable).
Then M ∼
= L∞ (X, µ) for some compact space X and probability measure µ on X.
This follows from the previous theorem by a result of von Neumann himself [23]:
Proposition 4.19 Let M ⊂ B(H) be an abelian von Neumann algebra (H separa-
ble). Then M = W ∗ (a) for some a = a∗ ∈ B(H) (i.e., M is “singly generated”).
The proof is a (difficult) exercise. Finally, the most general result is as follows:54
Theorem 4.20 Let M ⊂ B(H) be an abelian von Neumann algebra. Then one has
M∼= L∞ (X, µ) for some locally compact space X and Borel measure µ on X.
Conversely, L∞ (X, µ) defines a von Neumann algebra of multiplication operators on
H = L2 (X, µ), with operator norm equal to the norm k · kess,µ
∞ [8, Thm. ii.1.5], so
that we have found a complete characterization of abelian von Neumann algebras!
51
Of course, M = C falls in both classes, but is unique as such.
52
Here L∞ (σ(a), µ) denotes the space of (equivalence classes of) Borel functions f : σ(a) → C for
which f is bounded on σ(a)\∆0 for some Borel subset ∆0 ⊂ σ(a) with µ(∆0 ) = 0. The canonical
norm on L∞ (σ(a), µ) is kf kess,µ
∞ = inf{sup{|f (x)|, x ∈ σ(a)\∆}∆ ⊂ σ(a), | µ(∆) = 0}.
53
For the existence of µ see [8, §ix.8] for separable H, and [10, §i.7.2] in general.
54
The proof is technical (cf. [10, Thm. i.7.3.1] or [33, Vol. I, §iii.1]), but the idea is to find an
abelian C ∗ -algebra A for which M = A00 , upon which X = Σ(A), and the measure µ is constructed
such that µ(∆) = 0 iff µψ (∆) = 0 for all unit vectors Ψ ∈ H, with µψ defined similarly to (4.35).
One cannot take A = M , since Σ(M ) may not support such measures, cf. [33, Vol. I, Thm. iii.1.18].
4 VON NEUMANN ALGEBRAS 44
55
Under special assumptions, at least some good models for (X, µ) obtain. To explain the two
main examples, let us call a projection p ∈ P(M ) minimal if p 6= 0 and there exists no q ∈ P(M )
such that 0 < q < p (where q < p iff q 6 p and q 6= p). Then M is called atomless if it has no
minimal projections, whereas it is said to be atomic if for any nonzero p ∈ P(M ) there is a minimal
projection q ∈ P(M ) such that q 6 p. Under the assumption that H is separable, we then have
M∼ = L∞ (0, 1) (w.r.t. Lebesgue measure) iff M is atomless [33, Vol. I, Thm. iii.1.22], whereas if M
is atomic, it is isomorphic to either `∞ or to a finite direct sum of copies of C (exercise).
56
Equivalently, a Stone space is compact, T0 , and has a basis of clopen sets.
57
Recall that a lattice L is called distributive when x ∨ (y ∧ z) = (x ∨ y) ∧ (x ∨ z), and ortho-
complemented when there exists a map ⊥ : L → L that satisfies x⊥⊥ = x, y ⊥ 6 x⊥ when x 6 y,
x ∧ x⊥ = 0, and x ∨ x⊥ = 1. For example, P(M ) is orthocomplemented by p⊥ = 1 − p.
4 VON NEUMANN ALGEBRAS 45
Note that CG has a unit, namely δe (where δP e (x) = δe,x ). Next, represent CG on the
Hilbert space H = `2 (G) by f 7→ πL (f ) = x f (x)UL (x), where the (left) regular
representation UL of G is defined by UL (x)ψ(y) = ψ(x−1 y). The von Neumann
algebra of G is defined as W ∗ (G) = πL (CG)00 , so that W ∗ (G) ⊂ B(`2 (G)).
To see what is going on here, consider the case where G is finite,60 with an asso-
ciated finite set Ĝ of (unitary) equivalence classes of irreducible representations.61
∼ dγ
L γ ∈ Ĝ, realized on Hγ = C (with dγ < ∞),
Take some representative Uγ of each
and define a Hilbert space HĜ = γ∈Ĝ Mdγ (C) consisting of matrix-valued functions
ψ̂ on Ĝ, ψ̂(γ) ∈ Mdγ (C), with inner product (ψ̂, ϕ̂) = γ Tr (ψ̂(γ)∗ ϕ̂(γ)). Then UL
P
Indeed, writing ÛL (x) ≡ uUL (x)u−1 , we easily find ÛL (x)ψ̂(γ) = Uγ (x)ψ̂(γ), and
similarly π̂L (f )ψ(γ) = (1/ dγ )fˆ(γ)ψ(γ), where π̂L (f ) = uπL (f )u−1 . Using Schur’s
p
lemma (i.e., πγ (G)00 = Mdγ (C)), we obtain uW ∗ (G)u−1 = HĜ , seen as a ∗ -algebra,
acting on itself (now in its original guise as a Hilbert space) by fˆψ̂(γ) = fˆ(γ)ψ̂(γ).
Thus W (G) ∼
∗
L
= γ∈Ĝ Mdγ (C) is isomorphic to a finite direct sum of matrix algebras,
whose center W ∗ (G)∩W ∗ (G)0 is isomorphic to C(Ĝ). InPfact, ˜
p f : Ĝ → C corresponds
to ψ̂(γ) = f˜(γ)·1dγ in HĜ and hence to f (x) = (1/|G|) γ dγ f˜(γ)χγ (x) in W ∗ (G).
In particular, f lies in the center of W ∗ (G) iff it is a class function. This also
follows from (4.41), since g ∗ g = g ∗ f for all g iff f (yxy −1 ) = f (x) for all x, y ∈ G.
60
See, for example, [31] for the finite group theory used here (all based on the Peter–Weyl theory).
61
Recall that the cardinality of Ĝ equals the number kG of conjugacy classes in G. This is
because the class functions on G by definition form a kG -dimensional subspace C(G)c of `2 (G),
whilst the characters x 7→ χγ (x) = Tr Uγ (x), γ ∈ Ĝ, form a basis of C(G)c .
62
Unitarity follows from the fact that the matrix elements x 7→ (ei , Uγ (x)ej ), where (ei ) is an
o.n.b. of Hγ , form a basis of `2 (G), combined with Schur’s orthogonality relations. This implies in
turn that each irreducible representation Uγ of G, is contained in UL with multiplicity dγ .
4 VON NEUMANN ALGEBRAS 47
Lemma 4.35 For any a ∈ B(H) there exists √ a unique partial isometry w
∗
such that a = w|a| = |a |w, where |a| = a∗ a (defined by the continuous
functional calculus), the initial projection w∗ w is [(ker a)⊥ ] = [ran(a∗ )], and
the final projection ww∗ is [ran(a)].
But by Theorem 4.38, the right-hand side of (4.47) must be of the form diag(b, b) for
(1) (2)
some b ∈ Mat2 (M ), so that α̃t (a) = α̃t (a). This allows us to replace ∆it4 a22 ∆−it
4
in (4.48) by ∆it2 a22 ∆−it
2 . We then put U t = ∆it −it
1 ∆4 , which, unlike either ∆it
1 or
−it (1)
∆4 , lies in M , because each entry in α̃t (a) must lie in M if all the aij do, and
here we have taken a12 = 1. All claims of the theorem may then be verified using
elementary computations with 2 × 2 matrices. For example, combining
a 0 0 1 0 0 0 0
=
0 0 0 0 0 a 1 0
(1) (1) (1)
with the property α̃t (ab) = α̃t (a)α̃t (b), we recover (4.45). Using the identity
0 Ut 0 1 0 0
=
0 0 0 0 0 Ut ,
and evolving each side to time s, we arrive at (4.46). A ‘Proof from the Book’ !
4 VON NEUMANN ALGEBRAS 54
The Connes spectrum Γ(α) is a closed subgroup of R+ ∗ , which has the great virtue
0 0
that if π(α(R)) = π(α (R)), then Γ(α) = Γ(α ). So if α is the modular group of
M with respect to some state ω, then Γ(α) is independent of ω and may therefore
be called Γ(M ). This invariant can also be defined through the usual spectrum of
self-adjoint operators on Hilbert space. To this effect, Connes defined and proved
\ \
S(M ) := σ(∆ω ) = σ(∆ϕe ), (4.50)
ω 06=e∈P (M α )
where the first intersection is over all σ-weakly continuous faithful states ω on M ,
whereas in the second one takes a fixed σ-weakly continuous faithful state ϕ on M ,
and restricts it to ϕe := ϕ|Me . Furthermore, ∆ω denotes the operator ∆ on Hω ,
defined w.r.t. the usual cyclic unit vector Ωω of the gns-construction, etc. If M is
a type i or ii factor (on a separable Hilbert space) one has S(M ) = {1}, whereas
0 ∈ S(M ) iff M is type iii. Connes showed that Γ(M ) = S(M ) ∩ R+ ∗ , and the known
classification of closed subgroups of R+
∗ yields his parametrization of type iii factors:
Definition 4.41 Let M be a type iii factor. Then M is said to be of type:
• iii0 if Γ(M ) = {1};
• iiiλ , where λ ∈ (0, 1), if Γ(M ) = λZ ;
• iii1 if Γ(M ) = R+
∗.
4 VON NEUMANN ALGEBRAS 55
such that u(ei ⊗ ψ) = ui ψ for all ψ ∈ H 0 and i ∈ I, and finally prove that
u∗ M u = B(K) ⊗ 1H 0 .
67
If you like, submit nos. 5 and 6 by mail to A.J. Lindenhovius, imapp, fnwi, ru, Heyen-
daalseweg 135, 6525 aj Nijmegen, in order to replace your lowest mark for the earlier weeks.
4 VON NEUMANN ALGEBRAS 56
(a) Pick a nonzero finite projection p in M and show that thereP exists an
orthogonal family (pi )i∈N of projections with pi ∼ p for all i and i pi = 1.
In what follows, ui is a partial isometry in M such that u∗i ui = p and
ui u∗i = pi .
(b) Let H 0 = pH and consider N = pM p as a von Neumann algebra on
B(H 0 ). Prove that N is a factor, type ii1 factor.
(c) Show that the operator u : H 0 ⊗ `2 → H defined by u(ψ ⊗ ei ) = ui ψ is
unitary and satisfies uN ⊗ B(`2 )u∗ = M .
4 VON NEUMANN ALGEBRAS 57
3. The Hilbert–Schmidt operators B2 (H) form a Hilbert space in the inner product
Proof.
1. Although kbk ≥ kbvk for all b ∈ B(H) and all unit vectors v, for every > 0
there is a v ∈ H of norm 1 such that kbk2 ≤ kbvk2 + . Put b = (a∗ a)1/4 , and
note that k(a∗ a)1/4 k2 = kak. Completing v to a basis {ei }i , we have
X
kak = k(a∗ a)1/4 k2 ≤ k(a∗ a)1/4 vk2 + ≤ k(a∗ a)1/4 ei k2 + = kak1 + .
i
Similarly, kak ≤ |ak2 . The remaining inequality in (4.61) will follow from the
next item.
4 VON NEUMANN ALGEBRAS 59
P
2. Let aP ∈ B1 (H). Since i (ei , |a|ei ) < ∞, for every > 0 we can find n such
that i>n (ei , |a|ei ) < . Let pn be the projection onto the linear span of
{ei }i=1,...,n . Using (4.66), we have
kp⊥
n |a| k = kp⊥
1/2 2 ⊥ ⊥ ⊥
n |a|pn k ≤ kpn |a|pn k1 < .
Since p⊥
n = 1 − pn , it follows that pn |a|
1/2
→ |a|1/2 in the norm topology. Using
continuity of the involution and of multiplication, it follows that pn |a|pn → |a|
in norm. Since each operator pn |a|pn obviously has finite rank, |a| is compact.
But a has polar decomposition a = u|a|, so that a is compact (since B0 (H) is
a two-sided ideal in B(H)). The proof for B2 (H) is analogous.
P p
From the spectral theorem for compact operators, one now has kakp = λi
for p = 1, 2, where the λi ≥ 0 are the eigenvalues of |a| ≥ 0. The second
inequality in (4.61) is now immediate. Hence (4.61) has been proved.
3. The polarization formula (a + b)∗ (a + b) + (a − b)∗ (a − b) = 2(a∗ a + b∗ b)
yields the inequality (a + b)∗ (a + b) ≤ 2(a∗ a + b∗ b), since c∗ c ≥ 0 for all c,
including c = a − b. Hence a, b ∈ B2 (H) implies a + b ∈ B2 (H). With the
obvious λa ∈ B2 (H) when a ∈ B2 (H) for all λ ∈ C, this proves that B2 (H)
is a vector space. The sesquilinear form (4.63) is clearly positive semidefinite,
so the Cauchy–Schwarz inequality holds, and B2 (H) is a pre-Hilbert space.
In fact, by (4.61) the form (4.63) is positive definite, since k · k is a norm.
Finally, we show that B2 (H) is complete. Pick a basis {ei }i∈I in H (not
necessarily countable), and note
P that B2 (H) is the closure of the linear span of
all operators of the form a = i,j aij |ei ihej |. This is because of the continuity
of the inclusions in (4.62) and the fact that B0 (H) is itself the closure of this
linear span. An easy calculation gives
X X
k aij |ei ihej |k22 = |aij |2 .
i,j i,j
Applying (4.64) with ku∗ k ≤ 1, one has ka + bk1 ≤ kak1 + kbk1 . Hence B1 (H)
is a vector space and k · k1 is a norm; it is positive definite by (4.61). We
now prove completeness of B1 (H). Let {an } be a Cauchy sequence in k · k1 .
By (4.61), this is also a Cauchy sequence for k · k, which converges to some
a ∈ B0 (H). Let a − an = u|a − an |, so that |a − an | = u∗ (a − an ). Writing pN
for the projection on the linear span of e1 , . . . , eN , one has, for N < ∞,
N
X
(ei , |a − an |ei ) = lim Tr (pN u∗ (am − an )) ≤ lim sup kam − an k1 ,
m→∞ m
i=1
where we used (4.64) to derive the inequality. Since the right-hand side is
independent of N , we can let N → ∞ on the left, to obtain
8. It is clear from (4.64) that Tr (a) < ∞. To prove absolute convergence, with
a = u|a|1/2 |a|1/2 one has
Taking the sum and using the Cauchy–Schwarz inequality, one obtains
X
|(ei , aei )| ≤ k|a|1/2 u∗ k2 k|a|1/2 k2 .
i
By the argument in the proof of (4.64) (with b replaced by a), the right-hand
side is finite. Independence of the basis is now an easy exercise.
2. For part 2 of the theorem we also need this equality for ρ ∈ B0 (H)∗ . To that
effect, and replace the above estimate by
3. Now let a ∈ B(H) and ã ∈ B1 (H)∗ . From (4.64) and (4.65) one has kãk ≤ kak.
For the converse, take v, w ∈ H. On the one hand, one has
|ã(|wihv|)| ≤ kãkk|wihv|k1 .
Appendix 3
Half of Theorem 4.9 evidently follows from Theorem 4.10. The converse (‘if’) im-
plication uses a refinement of the gns-construction, where the state ω is assumed
to be σ-weakly continuous (such states are also called normal). In that case, using
the theory of σ-weakly closed ideals of von Neumann algebras, it can be shown that
πω (M ) coincides with πω (M )00 and hence is a von Neumann algebra [1, Thm. 2.4.24].
Since normal pure state on a von Neumann algebra may not exist, the ‘crazy’
Hilbert space Hc in the proof ofL Theorem 3.20 (see §3.10) must be replaced by the
even crazier direct sum Hec = ω∈Sn (M ) Hω , where this time the sum is over all
normal states on M . Similarly, in Lemma 3.8 one should now have a normal state
instead of a pure state. Otherwise, the proof that M has a faithful representation
as a von Neumann algebra on a Hilbert space is essentially follows the proof of
Theorem 3.20 (see Sakai’s own book [30, Thm. 1.16.7] for details).
We next prove (4.37). The inclusion M ⊂ M ⊥⊥ is trivial. For the converse, pick
a∈/ M ; since M is a von Neumann algebra, it is σ-weakly closed, so its complement
M c in B(H) is σ-weakly open. Hence there are ϕ ∈ B(H)∗ and > 0 such that the
open neighbourhood
kϕ̇k := inf{kϕ + ψk | ψ ∈ M ⊥ }.
where ϕ̇ is the image of ϕ ∈ B(H)∗ under the canonical projection, and the norm
is the one in B(H)∗ . Let ϕ := ϕ M be the restriction of ϕ ∈ B(H)∗ to M . It
is clear that the map ϕ 7→ ϕ̇ is well defined and is a linear bijection from M∗ to
B(H)∗ /M ⊥ . In fact, this map is isometric. Firstly, one trivially has
since now the supremum is taken over a larger set. Hence kϕ k ≤ kϕ̇k. Conversely,
for any ϕ ∈ B(H)∗ with kϕ̇k = 1 there exists, by a version of the Hahn–Banach
theorem, an a ∈ B(H) with â ∈ M ⊥⊥ , ϕ(a) = 1 and kak = 1. From (4.37) one then
infers that kϕ k ≥ |ϕ(a)| = 1 = kϕ̇k. This finishes the proof of Theorem 4.10.
REFERENCES 63
References
[1] Bratteli, O. & Robinson, D.W. (1987). Operator Algebras and Quantum Statisti-
cal Mechanics. Vol. I: C ∗ - and W ∗ -Algebras, Symmetry Groups, Decomposition
of States. 2nd Ed. Berlin: Springer.
[2] Bratteli, O. & Robinson, D.W. (1981). Operator Algebras and Quantum Sta-
tistical Mechanics. Vol. II: Equilibrium States, Models in Statistical Mechanics.
Berlin: Springer.
[3] A. Connes, Une classification des facteurs de type iii, Ann. Sci. École Norm.
Sup. 6, 133–252 (1973).
[4] A. Connes, Classification of injective factors. Cases ii1 , ii∞ , iiiλ , λ 6= 1, Ann.
Math. 104, 73–115 (1976).
[8] J.B. Conway, A Course in Functional Analysis (Springer, New York, 1990).
[11] R. Dudely, Real Analysis and Probability (Wadsworth & Brooks/Cole, Pacific
Grove, 1989).
[12] E.G. Effros and Z.-J. Ruan, Operator spaces (Oxford University Press, New
York, 2000).
[16] R.V. Kadison and J.R. Ringrose, Fundamentals of the Theory of Operator Al-
gebras, Vol. I: Elementary Theory (Academic Press, New York, 1983).
[17] R.V. Kadison and J.R. Ringrose, Fundamentals of the Theory of Operator Al-
gebras, Vol. II: Advanced Theory (Academic Press, New York, 1986).
REFERENCES 64
[18] N.P. Landsman and Ch.G. van Weert, Real- and imaginary-time field theory
at finite temperature and density, Phys. Rep. 145, 141–249 (1987).
[19] J. Lurie, Math261y: von Neumann algebras (Harvard University), Lecture notes
(in progress) available at www.math.harvard.edu/∼lurie/261y.html.
[20] F. J. Murray and J. von Neumann, On rings of operators, Ann. Math. 37,
116–229 (1936).
[21] F. J. Murray and J. von Neumann, On rings of operators, ii, Trans. Amer.
Math. Soc. 41, 208–248 (1936).
[22] F. J. Murray and J. von Neumann, On rings of operators, iv, Ann. Math. 44,
716–808 (1943).
[23] J. von Neumann, Zur Algebra der Funktionaloperatoren und theorie der nor-
malen Operatoren, Math. Ann. 102, 370–427 (1929).
[25] J. von Neumann, On rings of operators, iii, Ann. Math. 41, 94–161 (1940).
[26] J. von Neumann, On rings of operators. Reduction theory, Ann. Math. 50,
401–485 (1949).
[27] J. von Neumann, Collected Works. Vol. III: Rings of Operators, ed. A. H. Taub
(Pergamon Press, New York, 1961).
[28] G.K. Pedersen, Analysis Now, 2nd printing (Springer, New York, 1995).
[29] M. Reed and B. Simon, Methods of Modern Mathematical Physics, Vol. I: Func-
tional Analysis (Academic Press, San Diego, 1980).
[32] M. Takesaki, Twenty-five years in the theory of type iii von Neumann algebras,
in: R.S. Doran (ed.), C ∗ -algebras: 1943–1993, A Fifty Year Celebration, pp.
233–239 (AMS, Providence, 1994).
[35] S. Wright, Uniqueness of the injective iii1 factor (Lecture Notes in Mathematics
1413, Springer-Verlag, Berlin, 1989).