3.1 The last part of the proof
We can now return to Clifford algebras and Bott periodicity. The idea of Atiyah and Singer [3] was to extend the Atiyah–Jänich theorem to provide explicit descriptions of the loop spaces ΩkF(H). It turns out we can interpret this space (up to homotopy equivalence, of course) as Fredholm homomorphisms of Clifford algebra representations.
Note that Ck naturally comes with an involution defined by ei∗=−ei, and we shall require our representations to respect this. Let H=H0⊕H1 be a Z2-graded Hilbert space with a simultaneous graded action of all Ck, i.e. there are bounded linear operators J1,J2,… of degree 1 such that
JiJj+JjJi=0,Ji2=−1,Ji∗=−Ji,
for all i=j. We will focus on skew-Hermitian operators. Since we have a graded Hilbert space, we require our skew-Hermitian operators to have degree 1. So that B takes the form
(0T−T∗0).
Definition
We let
F^(H)Fk(H)={B∈F(H):B is skew-Hermitian}={B∈F^:BJi+JiB=0 for i=1,…,k}
After picking an arbitrary isomorphism
H0≅H1 of vector spaces, sending
B∈F^(H)=F0(H) to
T∈F(H0) as above gives a homeomorphism
F(H0)→F0(H). Hence
F0 represents
KO=KO0.
It turns out these Fk(H) are not what we want. To understand this, suppose B∈Fk(H) is unitary (which we can achieve by restricting to kerB⊥ and then rescaling), so that B2=−B∗B=−1. Then B “acts as” Jk+1 on H, and this gives H a new Ck+1 action. If k≡−1(mod4), then there is a unique irreducible Ck+1 module, hence the structure of H as a Ck+1-module is completely determined. However, if k≡−1(mod4), then we want to ensure H has infinitely many copies of each irreducible module, for it to be well-behaved.
Recall that to count how many copies of each each irrep we've got, we are supposed to count the eigenvalues of w=e1e2⋯ek after turning H0 to a Ck-module. The process of turning H0 into a Ck-module involves the inclusion Ck↪Ck+1, which sends ei to eiek+1. In our case, ek+1=B, and ei=Ji for i=1,…,k. Thus, we are counting the eigenvalues of
w=(J1B)(J2B)(J3B)⋯(JkB).
Since k≡−1(mod4) and B2=−1, this is equal to
w=J1J2⋯JkB.
We then want to require that the eigenvalues ±1 to both have infinite multiplicity. In general, for an arbitrary B, it is not unitary, or even injective. The desired “niceness” property is now that when we restrict B to the kerB⊥, and then rescale B so that it is unitary, the multiplicity of ±1 in w are both infinite. Equivalently, without doing the restriction and rescaling business, we want J1J2⋯JkB to have infinitely many eigenvalues of each sign, counted with multiplicity.
Definition
We say a self-adjoint operator is essentially definite if all but finitely many of its eigenvalues are of the same sign.
Definition
If k≡−1(mod4), we define F∗k(H)=Fk(H). Otherwise, we define
F∗k(H)={B∈Fk(H):J1J2⋯JkB∣H0 is not essentially definite}.
Observe that these form a component of
Fk(H).
The main theorem of Atiyah and Singer is the following:
Theorem
A
For k≥1, there is a homotopy equivalence
F∗k(H)→Ω(F∗k−1(H)).
Thus, combined with the Atiyah–Jänich theorem, we know that F∗k(H) represents KO−k.
We first see how this proves that KO−k(∗)=Ak, and in particular implies Bott periodicity.
Theorem
(Bott periodicity)
There is a natural homotopy equivalence F∗k+8≃F∗k.
Let
M be an irreducible
C8-module. Then there is an isomorphism
F∗k(H)A≅F∗k+8(H⊗^M)↦A⊗I,
and there is a contractible space of Ck+8-module isomorphisms H≅H⊗^M.
Theorem
(Computation of KO−k(∗))
The map idx:F∗k(H)→Ak defined by A↦[kerA] is continuous, and induces a bijection π0(F∗k)→Ak.
This in fact gives us an isomorphism of
rings.
For convenience, we drop the
(H) in
F∗k(H) when it is clear.
We leave the k=0 case for the reader. Note that C0=R is concentrated in degree 0, and it has two irreducible graded representations given by 0⊕R and R⊕0.
For k>0, we first show that the kernel map is continuous, so that it factors through π0(F∗k). Let B∈F∗k. Note that B2 is self-adjoint and negative, since
⟨B2x,x⟩=−⟨Bx,Bx⟩≤0.
Hence we know that σ(B2)⊆(−∞,0]. In fact, since B is Fredholm, we know 0 is an isolated point in the spectrum, and by scaling B, we may assume
σ(B2)⊆(−∞,−1]∪{0}.
Since the spectrum depends continuously on B2, we can pick a small neighbourhood of B in F^∗k such that whenever C is in the neighbourhood,
σ(C2)⊆(−∞,−1+ε]∪[−ε,0],
and further that ∥B2−C2∥<ε<21.
Let E be the spectral projection to [−ε,0]. We claim that E is isomorphic to kerB. Indeed, consider the orthogonal projection P from E to kerB.
P is injective. If not, suppose x∈E∩(kerB)⊥, and wlog assume ∥x∥=1. Since σ(B2∣kerB⊥)⊆(−∞,−1], we have
∣⟨(B2−C2)x,x⟩∣=∣⟨B2x,x⟩−⟨C2x,x⟩∣≥1−ε,
a contradiction.
P is surjective. If not, suppose x∈kerB∩E⊥, and again assume ∥x∥=1. Then
⟨(B2−C2)x,x⟩=−⟨C2x,x⟩≥1−ε,
a contradiction.
Now B2 and C2 commute with Ck, so the orthogonal projection is in fact a Ck-module isomorphism. We write E=kerC⊕kerC⊥. Then
idxB−idxC=[kerC⊥].
But the restriction of C to kerC⊥ is non-singular and skew-Hermitian, so we can use the action of C to turn kerC⊥ into a Ck+1 module. Morally, we should be able to just “scale” C, and the correct thing to write down is
Jk+1=C(−C2)−1/2.
Surjectivity is trivial (for [M]∈Ak, pick an isomorphism H≅H⊕M and use Jk+1⊕0∈F∗k(H⊕M)).
Injectivity follows from the following sequence of observations:
Claim
Any B∈F∗k can be deformed so that it is unitary on the complement of kerB.
By restricting to the complement of
kerB, we may assume
kerB is trivial. Note that the operator
B(−B2)−1/2 is unitary. So we can use the path
B(t−(1−t)B2)−1/2.
Claim
Let B∈F∗k, and let V be a Ck+1 module. Then we can deform B to B′ so that kerB′=kerB⊕V as a Ck-module.
We may assume
B is unitary on the complement of
kerB. Since
B∈F∗k, we know each
Ck+1 module appears as a direct summand of
kerB⊥ when
B acts as
Jk+1. So we can decompose
H=kerB⊕V⊕remaining,
and then linearly scale B to vanish on V.
Thus, if B,C∈F∗k are such that [kerB]=[kerC], then we can deform B and C so that their kernels are in fact isomorphic.
Claim
If B,C∈F∗k are such that kerB≅kerC as Ck-modules, then there is a Ck-module isomorphism T:H→H such that TBT−1=C.
Again wlog assume
B and
C are unitary on the complement, so that they act as
Jk+1. Then the complements are isomorphic as
Ck+1 modules, since both have infinitely many copies of each irrep. Hence there must be a
Ck-module isomorphism that sends
B to
C.
We can then conclude the theorem if we can find a path from T to the identity.
Claim
The group of Ck-module isomorphisms H→H is connected.
By Schur's lemma, this group is isomorphic to
GL(H), hence is in fact contractible.