NcDiffGeom_SoSe19_Lecture
NcDiffGeom_SoSe19_Lecture
differential geometry
held by Dr. Tobias Mai in Summer ’19
Contents
1. Introduction 5
Appendix 43
3
Chapter 1.
Introduction
(i) Take a classical space, i.e., a set X endowed with some additional structure
(e.g. a topological- or measure space, groups, manifolds, Lie groups, . . . );
(ii) Consider a suitable algebra of functions on X (e.g., C0 (X), C(X), L∞ (X),
C ∞ (X), . . . );
(iii) Transfer the additional structure of the space X to its associated (com-
mutative) algebra of functions and provide an intrinsic characterisation of that
structure;
(iv) Drop the assumption of commutativitiy.
• spaces, that are “badly behaved” as point sets, but correspond naturally
to (noncommutative) algebras (e.g., Penrose tilings, the space of leaves of a
foliation, the phase space in quantum mechanics, . . . )
5
Chapter 1. Introduction
But even for classical situations that are purely commutative, this point
of view gives new insights. Within noncommutative differential geometry,
manifolds are studied by some spectral data. The following definition is at the
heart of that approach.
D : H ⊇ dom D −→ H,
such that for all elements a ∈ A the following holds: π(a) dom D ⊆ dom D and
the commutator [D, π(a)] := Dπ(a) − π(a)D extends to an operator in B(H).
γ : R −→ T, t 7−→ exp(it).
1
D0 : H ⊇ dom D0 −→ H, g 7−→ g 0
i
on dom D0 := C 1 (T), which is a symmetric operator. One can show that D0
is essentially self-adjoint; let D be its closure, which is thus selfadjoint. Then
6
Chapter 1. Introduction
then FDF −1 is the multiplication by (n)n∈Z , hence we see that for all λ ∈ C−Z
it holds (D − λ)−1 ∈ K(H); moreover, for f ∈ A and g ∈ C 1 (T), it holds
1
[D, π(f )]g = D0 (f g) − f D0 g = π(f 0 )g.
i
We will see, that more general manifolds M induce spectral triples in a similar
way. Much of the structure of M can be recovered:
Exercises
Exercise 1.1: Let (A, H, D) be a spectral triple and let V ∈ B(H) be any
selfadjoint operator. Prove that (A, H, DV ) for the unbounded operator DV
given by DV := D + V with domain dom(DV ) := dom(D) is again a spectral
triple.
7
8
classical space noncommutative counterpart reconstruction theory
locally compact (compact) Haus- (unital) C ∗ -algebras, noncommuta- Gelfand-Naimark theorem, FA I,
dorff topological space tive topology Corollary 10.17
compact Hausdorff topological space von Neumann algebras, noncommu- FA II, Theorem 8.15
with finite radon measure tative measure theory
compact topological group compact quantum group, “noncom- Tannaka-Krein theorem
mutative” group theory
compact oriented smooth manifold spectral triple (unbounded) K- Connes’ reconstruction theorem2
Chapter 1. Introduction
2
Connes, 1994
2
2013
Chapter 2.
Spectral triples are (supposed to be) the right framework to extend classical
differential geometry to the noncommutative world. It is however not clear
offhand, how usual manifolds fit into that frame. In this chapter, we will see
that indeed each compact oriented smooth manifold induces a commutative
spectral triple in a natural fashion.
(iv) An atlas A of M is called smooth if all its transition maps are smooth,
i.e., C ∞ . A chart (U, ϕ) is said to be smooth with respect to a smooth
atlas A, if A ∪ {(U, ϕ)} is again a smooth atlas. A smooth atlas A is
called maximal, if every chart (U, ϕ) that is smooth with respect to A
already belongs to A. Every smooth atlas A induces a maximal one by
9
Chapter 2. Spectral triples associated to manifolds
δ([f ]x0 · [g]x0 ) = δ([f ]x0 ) · g(x0 ) + f (x0 ) · δ([g]x0 ) ∀ [f ]x0 , [g]x0 ∈ Cx∞0 (M)
Next, we “glue” the tangent spaces, yielding the so-called tangent bundle.
10
Chapter 2. Spectral triples associated to manifolds
π −1 (Ui ∩ Uj )
τi τj
σi,j
(Ui ∩ Uj ) × Kn (x,v)7→σi,j (x,v)=(x,Si,j (x)v)
(Ui ∩ Uj ) × Kn
for a continuous map Si,j : Ui ∩ Uj → Gln (K) called the transition maps.
Definition 2.5 (Smooth vector bundle): Let M be a smooth manifold. An
n-dimensional smooth vector bundle over M is an n-dimensional topological
vector bundle over M, for which all transition maps are smooth.
Definition 2.6 (Tangent bundle): Let M be an n-dimensional smooth manifold
with maximal smooth atlas A = {(Ui , ϕi ) | i ∈ I}. We put T M := x∈M Tx M
`
11
Chapter 2. Spectral triples associated to manifolds
12
Chapter 2. Spectral triples associated to manifolds
as desired.
(3) By the above steps, we now have a well-defined map
(as f1 and f2 agree on U and ρ(x0 ) = 1) and thus D|x0 (f1 ) = D|x0 (f2 ), as
desired.
13
Chapter 2. Spectral triples associated to manifolds
where, for a (U, f ) representing [f ]x0 , ρ is any bump function for (U, x0 ) and
ρf ∈ C ∞ (M) is defined (in fact well-defined) by
0, if x ∈
/ supp(ρ),
(ρf )(x) =
ρ(x)f (x), if x ∈ U.
belongs to X(V ).
(8) Finally, we get a linear map
Ψ : der(C ∞ (V )) −→ X(V ),
Remark 2.10: Like vector spaces underly linear algebra, vector bundles underly
what can be seen as “parametrised” linear algebra. Indeed, various constructions
for vector spaces can be generalised to that setting.
Let X be a Hausdorff topological space and let E and F be vector bundles
over X (both real or complex) of dimension n and m, respectively.
(i) The Whitney sum (or direct sum) E ⊕ F is the vector bundle of dimension
n + m with (E ⊕ F )x = Ex ⊕ Fx for all x ∈ X.
(ii) The tensor product bundle E ⊗ F is the vector bundle of dimension n · m
with (E ⊗ F )x = Ex ⊗ Fx for all x ∈ X.
(iii) The homomorphisms bundle hom(E, F ) is the vector bundle of dimension
n · m with hom(E, F )x = hom(Ex , Fx ) for all x ∈ X.
Analogously, the dual bundle E ∗ (see assignment 2A, exercise 1 (i)), the exterior
product p E, the vector bundle multp (E) of all p-multilinear maps and the
V
vector bundles symp (E) and altp (E) of all symmetric– respectively alternating
p-multilinear maps can be defined.
14
Chapter 2. Spectral triples associated to manifolds
We say that (ρi )i∈I is subordinate to an open cover (Ui )i∈I of X, if supp(ρi ) ⊆ Ui
for all i ∈ I.
(iii) On a paracompact space X, each open cover (Ui )i∈I of X has a subor-
dinate partition of unity (ρi )i∈I . If X = M is a paracompact smooth manifold,
then each ρi can be chosen to be smooth.
(i) The dual bundle T ∗ M to the tangent bundle T M is called the cotangent
bundle.
15
Chapter 2. Spectral triples associated to manifolds
(iii) The exterior derivative is the unique family (dp )p≥0 of R-linear maps
dp : Ωp (M) → Ωp+1 (M) satisfiying
• For all f ∈ Ω0 (M) = C ∞ (M ) and all x ∈ M
(i) A smooth atlas is called oriented if all its transition maps are orientation
preserving.
(ii) We say M is orientable if it admits an oriented smooth atlas.
(iii) An orientation on M is a maximal oriented smooth atlas.
16
Chapter 2. Spectral triples associated to manifolds
on the subspace Cc∞ (M) ⊆ C ∞ (M) of compactly supported functions such that
the following condition is satisfied: For every local chart (U, ϕ) in the maximal
oriented smooth atlas A of M and for each f ∈ Cc∞ (M) with supp(f ) ⊆ U we
have that ˆ ˆ
f= (f ◦ ϕ−1 )(det[gk,l ]1≤k,l≤n )1/2 dλn . (2.1)
M ϕ(U )
Here, λ is the Lebesgue measure on Rn and the functions gk,l ∈ C ∞ (ϕ(U )) are
n
Note that {∂k |ϕ(x) | 1 ≤ k ≤ n} is the basis of Tϕ(x) Rn introduced in Exercise 2.1
(ii) and (dϕ)(x) : Tx M → Tϕ(x) Rn is the differential of ϕ at x, which is defined
∞
for δ ∈ Tx M and [f ]ϕ(x) ∈ Cϕ(x) (Rn ) by
Remark 2.18: (i) The assumption that M is oriented guarantees that the
right-hand side of Eq. (2.1) is well-defined, i.e., independent of the particular
choice of the chart (U, ϕ).
Using a partition of unity subordinate to the family (Ui )i∈I for a maximal
oriented smooth atlas A = {(Ui , ϕi ) | i ∈ I}, say (ρi )i∈I , one can then define
for general f ∈ Cc∞ (M)
ˆ ˆ
X
f := (ρi f ).
M i∈I M
which requires only that M is oriented and on the other hand the volume
form dvol ∈ Ωn (M) of an oriented Riemannian manifold (M, g); in general,
ω ∈ Ωn (M) is called a volume-form if ω vanishes nowhere, and a paracompact
smooth manifold M is orientable if and only if a volume form exists; in fact,
fixing an equivalence class of volume forms specifies an orientation and vice versa;
17
Chapter 2. Spectral triples associated to manifolds
dvol is chosen such that dvol(x), for each x ∈ M, is normalised with respect to
the inner product on n Tx∗ M induced by g, i.e., hdvol(x), dvol(x)iVn Tx∗ M = 1.
V
In the case p = n, h·, ·iVp Tx∗ (M) was used in (iii). The latter extend naturally
to inner products
18
Chapter 2. Spectral triples associated to manifolds
n
X
h(ω0 , . . . , ωn ), (η0 , . . . , ηn )i := hωp , ηp iΩpC (M)
p=0
Then D0 is essentially self-adjoint; let D be its closure, which we call the Hodge-
de Rham operator. The Hodge-de Rham triple (A, H, D) is a commutative
spectral triple in the sense of Definition 1.1. We call ∆ := D2 the Hodge
Laplacian.
Definition 2.20: Let V be a K vector space with an inner product h·, ·i. Put
V• Vp
K V :=
L
p≥0 K V . Then
• p
(−1)k+1 hvk , viv1 ∧· · ·∧vbk ∧· · ·∧vp .
^ X
⌞ : V ×V −→ V, v⌞(v1 ∧· · ·∧vp ) :=
K k=1
Remark 2.21: The proof of Theorem 2.19 relies mostly on techniques that are
(not yet) at our disposal. We can understand, however, how commutators
[D, φ(f )] for f ∈ A look on H. They are given by the Clifford multiplication
with df from the left, i.e., for all ω ∈ Ω•C we have
19
Chapter 2. Spectral triples associated to manifolds
by
p
(−1)k+1 hvk , viC v1 ∧ · · · ∧ vbk ∧ · · · ∧ vp .
X
v ⌞ (v1 ∧ · · · ∧ vp ) :=
k=0
d∗ (f ω) = f d∗ ω − df ⌞ ω.
To see this, we take η ∈ Ω•C (M) and compute with respect to the inner
product h·, ·i := h·, ·iΩ•C (M) that
hdη, f ωi = hf dη, ωi
= hd(f η), ωi − hdf ∧ η, ωi
= hf η, d∗ ωi − hη, df ⌞ ωi = hη, f d∗ ω − df ⌞ ωi
[D, π(f )] = df • ω
That [D, π(f )] extends to bounded linear operator on H, will be discussed later.
Further, we note that Ω•C (M) ⊆ dom d∗ , which justifies that D0 = d + d∗ is
densely defined with dom D0 = Ω•C (M). This can be shown with the help of the
Hodge star operator ∗ : Ω•C (M) → Ω•C (M) which associates to each ω ∈ ΩkC (M)
the unique ∗ω ∈ Ωn−k n−k
C (M) such that for all x ∈ M and η ∈ ΩC (M)
20
Chapter 2. Spectral triples associated to manifolds
Hs := {ω ∈ H | (1 + ∆)s/2 ω ∈ H}
for s ≥ 0 and uses the Rellich Lemma to show that H1 ,→ H0 is compact, which
implies that
(1 + ∆)−1/2 : H = H0 −→ H1 ,−→ H0 = H
is compact and hence (D − i1)−1 is compact.
Remark 2.22: If the manifold M carries more structure (i.e., spinc -manifold),
there is another spectral triple (A, D, H) associated to M, with D being the
Dirac operator. We do not go into details here.
Exercises
Exercise 2.1: (i) Let x0 ∈ Rn be given. For j = 1, . . . , n, we define a linear
∂f
map ∂j |x0 : Cx∞0 (Rn ) → R by ∂j |x0 ([f ]x0 ) := (∂j f )(x0 ) = ∂x j
(x0 ) for every germ
∞ n
[f ]x0 ∈ Cx0 (R ). Prove that {∂j |x0 | j = 1, . . . , n} forms a basis of the tangent
space Tx0 Rn .
(ii) Let M be a n-dimensional smooth manifold with the maximal smooth
atlas A = {(Ui , φi ) | i ∈ I}. Show that for every i ∈ I and each x0 ∈ Ui , the
linear map
Θi,x0 : Rn −→ Tx0 M
that is defined by
n
v j ∂j (f ◦ φ−1
X
Θi,x0 (v) [f ]x0 := i ) (φi (x0 ))
j=1
21
Chapter 2. Spectral triples associated to manifolds
Exercise 2.3: Complete the proof of Theorem 2.9 of the lecture by proving the
following assertions for a smooth manifold M of dimension n and an open
subset V ⊆ M:
where ρ : M → [0, 1], for a chosen representative (U, f ) of the given germ [f ]x0 ,
is a bump function for (U, x0 ).
(ii) The induced linear map Ψ : der C ∞ (V ) → X(V ), D 7→ Ψ(D) satisfies
22
Chapter 3.
23
Chapter 3. The geodesic distance in noncommutative geometry
Remark 3.2: (i) On a connected Riemannian manifold (M, g), the geodesic
distance induces a metric
dg : M × M −→ [0, ∞), (x0 , x1 ) 7−→ dg (x0 , x1 )
called the Riemannian distance function. Note that if x0 6= x1 , the geodesic
distance dg (x0 , x1 ) is a positive number. Indeed, for γ ∈ Γ(x0 , x1 ) and a local
chart (U, ϕ) with x0 ∈ U and x1 ∈ / U , we have for all t ∈ [0, T ] the equality
gγ(t) (γ 0 (t), γ 0 (t)) = hG(γ(t))v 0 (t), v 0 (t)i
where G = (gk,l )1≤k,l≤n : ϕ(U ) −→ Mn (R) defined for all x ∈ U by
gk,l (ϕ(x)) := gx (dϕ)(x)−1 (∂k |ϕ(x) ), (dϕ)(x)−1 (∂l |ϕ(x) )
(see Theorem 2.17) and the smooth map v : [0, T ] → Rn , t 7→ ϕ(γ(t)), where
T ∈ [0, 1] is chosen such that γ([0, T ]) ⊂ U .
Take r > 0 such that cl(B(ϕ(x0 ), r)) ⊂ ϕ(U ) and V := ϕ−1 (B(ϕ(x0 ), r))
which is an open subset of U . We find δ ∈ (0, 1] such that for all y ∈ B(ϕ(x0 ), r)
and ξ ∈ Rn it holds
δkξk ≤ hG(y)ξ, ξi1/2 ≤ δ −1 kξk.
Thus
ˆ T0
L(γ) ≥ gγ(t) (γ 0 (t), γ 0 (t))1/2 dt
0
ˆ T0
= hG(γ(t))v 0 (t), v 0 (t)i1/2 dt
0
ˆ T0 ˆ T0
0
≥δ kv (t)k dt ≥ δ v 0 (t) dt = δkv(T 0 ) − ϕ(x0 )k
0 0
for every T 0 ∈ (0, T ] with γ([0, T 0 ]) ⊂ V . By enlarging T and taking the limit
in T 0 , we infer that L(γ) ≥ δr > 0 and thus we have dg (x0 , x1 ) ≥ δr > 0.
(ii) The topology on M induced by the metric dg agrees with the given
topology on M. This can be shown by arguments similar to (i). In fact, one
shows that for each x0 ∈ M and a local chart (U, ϕ) with x0 ∈ U , an open
neighbourhood V ⊆ U of x0 and δ ∈ (0, 1], it exists r > 0 such that for all
x ∈ V it holds
δ|ϕ(x) − ϕ(x0 )| ≤ dg (x, x0 ) ≤ δ −1 |ϕ(x) − ϕ(x0 )|,
and dg (x, x0 ) ≥ δr for all x ∈ M − V .
24
Chapter 3. The geodesic distance in noncommutative geometry
Our goal is to “dualise” the definition of the geodesic distance such that it
fits into the framework of spectral triples.
Theorem 3.3 (Musical isomorphisms): Let (M, g) be a Riemannian manifold.
Then g can be seen as a positive definite pairing on smooth vector fields, i.e., a
map
g : X(M) × X(M) −→ C ∞ (M)
which is C ∞ (M)-bilinear and satisfies g(X, X) ≥ 0 for all X ∈ X(M) and
g(X, X)(x) = 0 at x ∈ M if and only if X(x) = 0.
This induces an isomorphism (in fact, a C ∞ (M)-bimodule map)
[ : X(M) −→ Ω1 (M), X 7−→ X [ := g(X, ·).
Its inverse ] : Ω1 (M) → X(M), ω 7→ ω ] is determined by ω(X) = g(ω ] , X) for
all X ∈ X(M). The inner product on Ω1 (M) defined in (Remark 2.18) (iii)
satisfies ˆ
hω, ηiΩ1 (M) = g(ω ] , η ] )
M
for all ω, η ∈ Ω (M, since hω(x), η(x)iTx∗ (M) = gx (ω ] (x), η ] (x)) for each
1
x ∈ M.
25
Chapter 3. The geodesic distance in noncommutative geometry
with d0 f ∈ Ω1 (M) as defined in (Definition 2.14) (iii). We thus have that for
each x ∈ M and δ ∈ Tx M
|(f ◦ γ)0 (t)| ≤ gγ(t) (gradγ(t) f, gradγ(t) f )1/2 gγ(t) (γ 0 (t), γ 0 (t))1/2
≤ kgrad f k∞ gγ(t) (γ 0 (t), γ 0 (t))1/2 .
In summary, we get |f (x1 ) − f (x0 )| ≤ kgrad f k∞ L(γ). We infer from the latter
that
26
Chapter 3. The geodesic distance in noncommutative geometry
1
|fε (x1 ) − fε (x0 )| = |hε (x1 ) − hε (x0 )
1+ε
1
= |f0 (x1 ) − (f0 (x1 ) − hε (x1 )) + (hε (x0 ) − f0 (x0 ))|
1+ε
1
≥ |f0 (x1 )| − (|f0 (x1 ) − hε (x1 )| + |f0 (x0 ) − hε (x0 )|
1+ε
1 2ε
≥ dg (x0 , x1 ) − ,
1+ε 1+ε
which shows that for every ε > 0 it holds
dg (x0 , x1 ) 2ε
sup{|f (x1 ) − f (x0 )| | f ∈ C ∞ (M) : kgrad f k∞ ≤ 1} ≥ − .
1+ε 1+ε
We conclude by taking the limit ε ↓ 0.
a C ∗ -algebra and denote by S(A) the state space of A. We define for ϕ, ψ ∈ S(A)
by
In view of Remark 3.2 (iv), the following theorem says that the Hodge-de
Rham triple remembers the metric.
Theorem 3.8: Let (M, g) be a compact oriented Riemannian manifold and
(A, H, D) be the Hodge-de Rham triple for (M, g). Then the faithful representa-
tion π : A → B(H) extends to a faithful ∗ -representation π̂ : C(M, C) → B(H)
which induces an isometric ∗ -isomorphism
∼
=
πb : C(M, C) A := clk·k (A) ⊆ B(H).
27
Chapter 3. The geodesic distance in noncommutative geometry
• •
Tx∗ MC −→ Tx∗ MC ,
^ ^
c(v) : ω 7−→ v · ω
C C
We conclude that
k[D, π(f )]ωk Ω•C (M) ≤ maxk(df )(x)k Tx∗ MC · kωkΩ•C (M)
x∈M
Optimising ω, we get that k[D, π(f )]k = kgrad f k∞ and thus, by (Theorem
3.6) and (Exercise 4AB - 1(ii))
28
Chapter 3. The geodesic distance in noncommutative geometry
Exercises
Exercise 3.1: Let (A, H, D) be a spectral triple with the faithful ∗ -representation
π : A → B(H). Consider the state space S(A) for the associated C ∗ -algebra
A := clk·k (π(A)) ⊆ B(H). Prove the following assertions:
(i) If the image of the set {a ∈ A | k[D, π(a)]k ≤ 1} under the canonical
projection in the quotient Banach space A/C1 is a norm bounded set,
then the spectral distance satisfies dD (φ, ψ) < ∞ for all φ, ψ ∈ S(A) and
induces a metric dD : S(A) × S(A) → [0, ∞).
(ii) For all φ, ψ ∈ S(A), we have that
dD (φ, ψ) = sup{|ψ(π(a)) − φ(π(a))| | a = a∗ ∈ A : k[D, π(a)]k ≤ 1}.
Hint: To prove “≤”, establish first that the set {a ∈ A | k[D, π(a)]k ≤ 1}
is closed under the following maps: a 7→ ζa for each ζ ∈ C with |ζ| = 1,
a 7→ a∗ , a 7→ Re(a) = 12 (a + a∗ ), and a 7→ Im(a) = 2i1 (a − a∗ ).
Exercise 3.2: Let (A1 , H1 , D1 ) and (A2 , H2 , D2 ) be spectral triples with the
faithful ∗ -representations π1 : A1 → B(H1 ) and π2 : A2 → B(H2 ), respectively.
We call these two spectral triples equivalent, if there exists a ∗ -isomorphism
Φ : A1 → A2 and a unitary operator U : H1 → H2 such that for all a ∈ A1 it
holds
U π1 (a)U ∗ = π2 (Φ(a))
and U D1 U ∗ = D2 .
Show that in this case adU : B(H1 ) → B(H2 ), x 7→ U xU ∗ is an isometry
which satisfies adU (A1 ) = A2 , where A1 and A2 are the C ∗ -algebras associated
to A1 and A2 , respectively, and prove that ad∗U : S(A2 ) → S(A1 ), φ 7→ φ ◦ adU
defines an isometry for the spectral distances, i.e.,
dD1 (ad∗U φ, ad∗U ψ) = dD2 (φ, ψ) for all φ, ψ ∈ S(A2 ).
Exercise 3.3: Consider the complex unital ∗ -algebra A = C ⊕ C with entry-
wise operations. Let H1 and H2 be finite dimensional complex Hilbert spaces
and put H := H1 ⊕ H2 . Define the ∗ -homomorphism π : A → B(H) for all
a = (a1 , a2 ) ∈ A by !
a1 idH1 0
π(a) := .
0 a2 idH2
Further, take any linear operator M : H1 → H2 and consider the operator
!
0 M∗
D := .
M 0
29
Chapter 3. The geodesic distance in noncommutative geometry
(i) Verify that (A, H, D) is a spectral triple. Compute for all a = (a1 , a2 ) ∈ A
the commutator [D, π(a)] and show that its norm is given by
30
Chapter 4.
In (Chapter 3), we have seen that the geodesic distance on a connected, compact
and oriented Riemannian manifold can be recovered from its associated Hodge-
de Rham triple via Connes spectral distance.
In this chapter, we will discuss the noncommutative integral, by which
integration of (smooth) functions with respect to the Riemann-Lebesgue measure
on Riemannian manifolds as introduced in (Theorem 2.17) is generalised to the
framework of spectral triples.
Like in quantum mechanics, the underlying idea is that operators on a
separable complex Hilbert space H with dim H = ∞ take over the role of
complex variables, while selfadjoint operators on H correspond to real variables.
Remark 4.1: Recall (Theorem 9.8 from the Functional Analysis I lecture notes)
that T ∈ B(H) is compact if and only if T can be approximated in operator
norm on B(H) by finite rank operators; equivalently,
∀ ε > 0 ∃ V ⊆ H subspace, dim V < ∞ : kT |V ⊥ k < ε,
where T |V ⊥ : V ⊥ → H is the restriction of T to V ⊥ and k·k is the norm on
B(V ⊥ , H).
For a compact operator T ∈ B(H), we call the non-zero eigenvalues (µn (T ))n≥0
of |T | := (T ∗ T )1/2 , arragned in decreasing order and repeated according to
multiplicity, the characteristic values of T . Note that µn (T ) converges to 0 as
n → ∞. We have that for all n ∈ N0
µn (T ) = inf{kT − Sk | S ∈ B(H) : dim ran S ≤ n}
(4.1)
= inf{kT |V ⊥ k | V ⊆ H subspace, dim V = n}
and in particular µ0 (T ) = kT k.
31
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
Remark 4.3: Recall (Theorem 9.5 from the Functional Analysis I lecture notes)
that K(H) is a norm-closed two-sided ideal in B(H). It follows from Eq. (4.1)
that for all T ∈ K(H) and S ∈ B(H)
is finite for some (and in turn for each) orthonormal basis (ξk )k∈N0 of H. In
this case, the sum ∞ k=0 hT ξk , ξk i is absolutely convergent and its value Tr(T )
P
32
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
Remark 4.6: Note that σN (T ) is a partial sum in Eq. (4.2). For T ∈ I1 (H),
we find constants C, C 0 < ∞ such that for all N ∈ N
N −1
C
≤ C 0 log(N ).
X
σN (T ) = kT k +
n=1 n
Proof: Exercise!
33
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
and extends uniquely to a positive linear map Trω : L (1,∞) (H) → C, which for
all S ∈ B(H), T ∈ L (1,∞) (H) and U ∈ L 1 (H) ⊆ L (1,∞) (H) satisfies
and
Trω (U ) = 0. (4.4)
⊆ L 1 (H). We call Trω a Dixmier trace.
S
Note that α>1 Iα (H)
Proof: For every T ∈ L (1,∞) (H)+ , (γN (T ))N ∈N is bounded (see Remark 4.6);
hence, Trω (T ) is well-defined and clearly Trω (T ) ≥ 0 by (i). For T1 , T2 ∈
L (1,∞) (H)+ , it follows from Proposition 4.7 and property (i) of ω that
In summary, we get that Trω (T1 + T2 ) = Trω (T1 ) + Trω (T2 ). That for all λ ≥ 0
and T ∈ L (1,∞) (H)+ it holds Trω (λT ) = λ Trω (T ) is clear since ω is linear and
γN (λT ) = λγN (T ) for each N ∈ N.
The extension of Trω to L (1,∞) (H) uses that
1
This condition is called the scale invariance.
34
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
Let U ∈ B(H) be unitary. To prove Eq. (4.3), we note first that Eq. (4.1)
implies for n ∈ N0 and for all T ∈ K(H) that µn (U T U ∗ ) = µn (T ), which yields
by definition for N ∈ N that for all T ∈ K(H) it holds γN (U T U ∗ ) = γN (T )
and hence it holds for all T ∈ L (1,∞) (H) that Trω (U T U ∗ ) = Trω (T ). Because
L (1,∞) (H) is a two-sided ideal, this is equivalent to the statement that for all
T ∈ L (1,∞) (H) it holds Trω (U T ) = Trω (T U ).
Since every S ∈ B(H) is a linear combination of (in fact four) unitaries (see
the proof of Lemma 6.13 in the Functional Analysis II lecture notes), the latter
yields Eq. (4.3).
To verify Eq. (4.4), we take T ∈ L 1 (H) and without loss of generality, we
may assume that T ≥ 0. Note that (σN (T ))n∈N is bounded by kT k1 due to
Eq. (4.2), thus γN (T ) → 0 as N → ∞, so that by property (iii) we have
Trω (T ) = ω (γN (T ))N ∈N = lim γN (T ) = 0.
N →∞
Note that Iα (H) ⊆ L 1 (H) for all α > 1, since for each T ∈ Iα (H), we find
C < ∞ such that
∞ ∞
n−α < ∞.
X X
kT |1 = µn (T ) ≤ kT k + C
n=0 n=1
Remark 4.10: (i) The existence of (in fact infinitely many) linear maps
ω : `∞ (N, R) → R satisfying the conditions (i), (ii) and (iii) in Theorem 4.8
was proved by Dixmier in 1966: With the construction of Trω , he proved the
existence of singular traces on B(H) (i.e., traces that vanish on L 1 (H)) and
settled to the negative the question of the uniqueness of the trace on B(H).
35
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
σλ (T )
γ : [a, ∞) −→ R, λ 7−→
log(λ)
is a continuous and bounded function, its Cesàro mean with respect to the
Haar measure du
u
on the multiplicative group (0, ∞) is for each λ ∈ [0, ∞) given
by ˆ λ
1 du
τλ (T ) := γ(u)
log(λ) a u
and defines a function λ 7→ τλ (T ) in Cb ([a, ∞)). For T ∈ L (1,∞) (H)+ , let τ̇ (T )
be the class of λ 7→ τλ (T ) in the quotient C ∗ -algebra B := Cb ([a, ∞))/C0 ([0, ∞)).
One can show that τ̇ : L (1,∞) (H)+ → B extends to a positive linear map
τ̇ : L (1,∞) (H) → B with the property that for each S ∈ B(H) and for each
T ∈ L (1,∞) (H) it holds τ̇ (ST ) = τ̇ (T S).
For every state ω on B, one defines Trω : L (1,∞) (H) → C by
Trω (T ) := ω(τ̇ (T ))
for all T ∈ L (1,∞) (H), this map Trω then satisfies Eq. (4.3) and Eq. (4.4) in
Theorem 4.8. Moreover, we have that T ∈ L (1,∞) (H) is measurable if and only
if limλ→∞ τλ (T ) exists, in which case
T = lim τλ (T ).
λ→∞
N =1
36
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
it follows that L (p,∞) (H) = I1/p (H) for each p > 1. On the diagonal, one finds
the Schatten-ideals L p (H) := L (p,p) (H), where the interpolation norm k·k(p,p)
on L (p,p) (H ) is equivalent to the Schatten-p-norm
kT kp := Tr(|T |p )1/p
for T ∈ L p (H). A noncommutative integration theory for which L p (H) serves
as a analoge of the Lp -space in Lebesgue integration theory, was developed by
Segal in the fifties.
Definition 4.11: Let (A, H, D) be a spetrac triple. We say that (A, H, D) is
(i) p-summable if (1 + D2 )−1/2 ∈ L p (H),
(ii) (p, ∞)-summable if (1 + D2 )−1/2 ∈ L (p,∞) (H),
(iii) θ-summable if e−tD ∈ L 1 (H).
2
Example 4.12: Consider the spectral triple (A, H, D) from Example 1.2, where
A = C ∞ (T, C), H = L2 (T, m) and D was the closure of
1
D0 : H ⊃ dom D0 −→ H, g 7−→ g 0
i
with domain dom D0 := C 1 (T). Then the operator ∆ := D2 has spectrum
Sp(∆) = {|n|2 | n ∈ Z}. We conclude that (1 + ∆)−1/2 ∈ L (1,∞) (H), i.e.,
the spectral triple (A, H, D) is (1, ∞)-summable. Indeed, since the Fourier
transform F : L2 (T, m) → `2 (Z) is a unitary and FDF −1 = M(n)n∈Z , where
Mλ , for any sequence λ = (λn )n∈Z of complex numbers, is the closed operator
Mλ : `2 (Z) ⊇ dom Mλ −→ `2 (Z), (an )n∈Z 7−→ (λn an )n∈Z
with domain dom Mλ := {(an )n∈Z ∈ `2 (Z) | (λn an )n∈Z ∈ `2 (Z)} we conclude
that
F∆F −1 = M(|n|2 )n∈Z , F(1 + ∆)−1/2 F −1 = M(1+n2 )−2 )n∈Z ∈ B(`2 (Z))
and finally (1 + ∆)−1/2 ∈ L (1,∞) (H); in fact, we have
!
−1/2
1 1 1 1
µn ((1 + ∆) ) = 1, √ , √ , . . . , √ , √ ,... ,
n∈N0 2 2 1 + n2 1 + n2
so that
ffl γN ((1 + ∆)−1/2 ) → 2 as N → ∞ and hence even (1 + ∆)−1/2 ∈ M(H)
with (1 + ∆)−1/2 = 2.
37
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
Example 4.13: In the situation of Example 4.12, let P ∈ B(H) be the orthogo-
nal projection onto ker D = C1 ⊂ L2 (T, m). By Exercise 1B-1, (A, H, D̃) with
D̃ := DP = D + P gives another spectral triple. Note that ∆ ˜ := D̃2 = ∆ + P ,
˜ becomes invertible. We have ∆
so that ∆ ˜ −1/2 ∈ L (1,∞) (H) since
˜ −1/2 ))n∈N = 1, 1, 1, 1 , 1 , . . . , 1 , 1 , . . .
(µn (∆ 0
2 2 n n
P : Γ(M, E) −→ Γ(M, E)
∂ α1 ∂ αn
Aα (x)(−i)|α|
X
P = · · ·
|α|≤m
∂xα1 1 ∂xαnn
|α|=d |α|=d
38
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
|α|=d
Theorem 4.15 (Connes’ trace theorem, 1988): Let (M, g) be a compact Rie-
mannian manifold of dimension n. For P ∈ Ψ−n (M, E) with a complex vector
bundle E over M, the following statements hold:
39
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
(i) P extends to a bounded linear operator on the Hilbert space L2 (M, E),
which is obtained by completion of Γ(M, E) with respect to the inner
product given by
ˆ
hu1 , u2 i := u2 (x)∗ u1 (x) dvol(x) .
M
Exercises
Exercise 4.1: Let H be an infinite dimensional separable complex Hilbert space.
(i) Let T ∈ K(H) and N ∈ N be given. Prove the formula
n o
σN (T ) = inf kRk1 + N kSk R ∈ L 1 (H), S ∈ K(H) : T = R + S
for the value σN (T ) that was defined in Definition 4.5 of the lecture.
(ii) Like in Remark 4.10 (ii), we define for every T ∈ K(H) and each λ > 0
n o
σλ (T ) := inf kRk1 + λkSk R ∈ L 1 (H), S ∈ K(H) : T = R + S .
Due to (i), this interpolates the values σN (T ). Show that this interpolation
is in fact piecewise linear, i.e., prove that σλ (T ) = λkT k holds for every
λ ∈ [0, 1) and that
σN +λ (T ) = (1 − λ)σN (T ) + λσN +1 (T )
holds for each N ∈ N and every λ ∈ [0, 1).
Exercise 4.2: Let (A1 , H1 , D1 ) and (A2 , H2 , D2 ) be spectral triples with infinite
dimensional separable complex Hilbert spaces H1 , H2 and suppose that the
operator Γ1 ∈ B(H1 ) is a grading on (A1 , H1 , D1 ). Put
A := A1 ⊗C A2 , H := H1 ⊗C H2 , and D := D1 ⊗ idH2 +Γ1 ⊗ D2 .
Prove the following assertions:
(i) (A, H, D) is a spectral triple.
(ii) (A, H, D) is θ-summable2 whenever at least one of the spectral triples
(A1 , H1 , D1 ) and (A2 , H2 , D2 ) is θ-summable.
2
Recall from Definition 4.11 that a spectral triple (A, H, D) is said to be θ-summable if
2
e−tD ∈ L 1 (H) for each t > 0.
40
Chapter 4. The Riemann-Lebesgue measure in noncommutative geometry
(ii) For each u ∈ S(Rn ) and each multi-index α ∈ Nn0 , we have that
aα (−i)|α| ∂ α
X
P =
|α|≤m
for some integer m ≥ 0 and with coefficients aα ∈ C ∞ (Ω) for each |α| ≤ m.
Let
pP : Ω × Rn −→ R, aα ξ α
X
(x, ξ) 7−→
|α|≤m
be the complete symbol of P . Prove that for each u ∈ S(Rn ) and every point
x ∈ Rn ˆ
1
(P u|Ω )(x) = n/2
eihξ,xi pP (x, ξ)û(ξ) dλn (ξ) .
(2π) Rn
41
Appendix
43
Appendix A.
Let (H1 , h·, ·i) and (H2 , h·, ·i2 ) be complex Hilbert spaces; the norms induced
by the inner product are denoted by k · k1 and k · k2 , respectively.
T : H1 ⊇ dom T −→ H2
It is thus linear subspace of the Hilbert space H1 ⊕ H2 with the inner product
given by
45
III. The adjoint operator
(i) T is closable;
(ii) For every sequence (xn )n∈N in dom T which converges to 0 in H1 and for
which (T xn )n∈N converges in H2 to a point y ∈ H2 , we necessarily have
that y = 0;
(iii) cl(G(T )) ∩ ({0} × H2 ) = {(0, 0)}.
It is worthwhile to take a closer look on the proof that (iii) implies (i). It
follows from Lemma A.1 that if (iii) holds, then cl(G(T )) must be the graph
of an unbounded linear operator, say T : H1 ⊇ dom T → H2 . The operator
T is thus a closed extension of T ; in fact, it is the (unique) minimal closed
extension (i.e., for every other closed operator S that satisfies T ⊆ S, it follows
that T ⊆ S), called the closure of T . Furthermore, its domain dom T is the
closure of dom T with respect to the graph norm k · kT which is defined by
kxk2T := kxk21 + kT xk22 for each x ∈ dom T .
46
which is clearly a subspace of H2 . Since dom T is dense in H1 , φy for every
y ∈ dom T ∗ extends uniquely to a bounded linear functional Φy on H1 ; by the
Riesz representation theorem, the latter must be of the form Φy (x) = hx, T ∗ yi1
for all x ∈ H1 with a unique vector T ∗ y ∈ H1 . The assignment y 7→ T ∗ y is in
fact linear on dom T ∗ , so that this construction results in an unbounded linear
operator
T ∗ : H2 ⊇ dom T ∗ −→ H1 ,
called the adjoint of T .
Lemma A.4:
47
V. Resolvent set and spectrum
Suppose in addition that T is closed. Then the following statements are equiva-
lent:
(i) T is selfadjoint;
(ii) n+ (T ) = n− (T ) = 0;
(iii) ran(T + i) = ran(T − i) = H.
48
(i) If (T − λ1) : dom T → H is bijective for a λ ∈ C, then its inverse
(T − λ1)−1 is bounded.
(ii) The spectrum σ(T ) ⊆ C is closed.
(iii) If T is selfadjoint, then σ(T ) ⊆ R.
(iv) If T is symmetric and satisfies σ(T ) ⊆ R, then T is selfadjoint.
If h : R → R is measurable, then
ˆ
hh(T )x, yi = h(λ) dhE(λ)x, yi
R
49
Appendix B.
We recall some basic facts about the Fourier transform on Rn which can be
used for the solution of the exercises without proof.
where h·, ·i denotes the standard inner product on Rn , i.e., hξ, xi = nj=1 ξj xj for
P
51
Appendix C.
Solution (to Exercise 1.1): According to Definition 1.1, we have to check that
(i) DV is selfadjoint,
(ii) DV has compact resolvent,
(iii) For all a ∈ A, π(a) dom DV is contained in DV and [DV , π(a)] is bounded
on H.
Ad (i): This is a general fact: If T : H ⊇ dom T → H and S ∈ B(H),
(T + S)∗ = T ∗ + S ∗ . In particular, if T and S are selfadjoint, then T + S is
also selfadjoint.
To see this, note that
since x 7→ hSx, yi is bounded for every y ∈ H. Thus, for all x ∈ dom T and for
all y ∈ dom T ∗ it holds
i.e., (T + S)∗ = T ∗ + S ∗ .
Ad (ii): Take any λ ∈ C − σ(DV ), choose λ1 ∈ C − σ(D) and λ2 := λ − λ1 .
Then
53
Appendix C. Solutions to the exercises
i.e., we have
{∂j |x0 | 1 ≤ j ≤ n}
Pn
is R-linearly independent. Let therefore α1 , . . . , αn ∈ R with j=1 αj ∂j |x0 = 0
in Tx0 Rn . Then, for each [f ]x0 ∈ Cx∞0 (R), it holds
n n
∂f
αj ∂j |x0 ([f ]x0 ) = αj
X X
0= (x0 ).
j=1 j=1 ∂xj
54
Appendix C. Solutions to the exercises
and hence
n
X ∂f
δ([f ]x0 ) = δ([fj ]x0 ) (x0 ) ∈ Lin{∂j |x0 | 1 ≤ j ≤ n},
j=1 ∂xj
j=1
55
Appendix C. Solutions to the exercises
x∈X
LEE (Si,j
∗
(x)) = τj∗ |Ex∗ ◦ (τi∗ |Ex∗ )−1
= ((τj |Ex )0 ◦ Φ)−1 ◦ ((τi |Ex )0 ◦ Φ)
= Φ−1 ◦ ((τi |Ex ) ◦ (τj |Ex )−1 )0 ◦ Φ
= Φ−1 ◦ (LEE (Sj,i (x)))0 ◦ LEE ∗ (In )
∗
= Φ−1 ◦ LEE ∗ (Sj,i (x)> ) ◦ Φ = LEE (Sj,i (x)> )
where Si,j : Ui ∩ Uj → Gln (K) are the transition matrices for E, E is the
standard basis of Kn , E ∗ is the dual basis to E and we fix Φ that satisfies
Φ(ei ) = e∗i for i ∈ {1, . . . , n}.
∗
From the above calculation we conclude Si,j (x) = Sj,i (x)> = (Si,j (x)−1 )> . In
∗
particular, if E is smooth, then E is smooth as well.
56