the atiyah-singer index theorem what it is and why you ... › indextheory › asit3.pdf · the...

Post on 25-Jun-2020

17 Views

Category:

Documents

0 Downloads

Preview:

Click to see full reader

TRANSCRIPT

The Atiyah-Singer Index theoremwhat it is and why you should care

Rafe Mazzeo

October 10, 2002

Before we get started:

At one time, economic conditions caused the closing of

several small clothing mills in the English countryside. A

man from West Germany bought the buildings and

converted them into dog kennels for the convenience of

German tourists who liked to have their pets with them

while vacationing in England. One summer evening, a local

resident called to his wife to come out of the house.

Before we get started:

At one time, economic conditions caused the closing of

several small clothing mills in the English countryside. A

man from West Germany bought the buildings and

converted them into dog kennels for the convenience of

German tourists who liked to have their pets with them

while vacationing in England. One summer evening, a local

resident called to his wife to come out of the house.

”Just listen!”, he urged. ”The Mills Are Alive With the

Hounds of Munich!”

Background:

The Atiyah-Singer theorem provides a fundamental link

between differential geometry, partial differential equations,

differential topology, operator algebras, and has links to

many other fields, including number theory, etc.

Background:

The Atiyah-Singer theorem provides a fundamental link

between differential geometry, partial differential equations,

differential topology, operator algebras, and has links to

many other fields, including number theory, etc.

The purpose of this talk is to provide some sort of idea

about the general mathematical context of the index

theorem and what it actually says (particularly in special

cases), to indicate some of the various proofs, mention a

few applications, and to hint at some generalizations and

directions in the current research in index theory.

Outline:

1. Two simple examples

Outline:

1. Two simple examples

2. Fredholm operators

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

6. The Gauß-Bonnet and Hirzebruch signature theorems

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

6. The Gauß-Bonnet and Hirzebruch signature theorems

7. The first proofs

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

6. The Gauß-Bonnet and Hirzebruch signature theorems

7. The first proofs

8. The heat equation and the McKean-Singer argument

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

6. The Gauß-Bonnet and Hirzebruch signature theorems

7. The first proofs

8. The heat equation and the McKean-Singer argument

9. Heat asymptotics and Patodi’s miraculous cancellation

Outline:

1. Two simple examples

2. Fredholm operators

3. The Toeplitz index theorem

4. Elliptic operators on manifolds

5. Dirac operators

6. The Gauß-Bonnet and Hirzebruch signature theorems

7. The first proofs

8. The heat equation and the McKean-Singer argument

9. Heat asymptotics and Patodi’s miraculous cancellation

10. Getzler rescaling

and if there is time:

11. Application: obstruction to existence of positive scalar

curvature metrics

and if there is time:

11. Application: obstruction to existence of positive scalar

curvature metrics

12. Application: the dimension of moduli spaces

and if there is time:

11. Application: obstruction to existence of positive scalar

curvature metrics

12. Application: the dimension of moduli spaces

13. The Atiyah-Patodi-Singer index theorem

The simplest example:

Suppose

A : Rn −→ Rk

is a linear transformation. The rank + nullity theorem

states that

dim ran (A) + dimker(A) = n.

The simplest example:

Suppose

A : Rn −→ Rk

is a linear transformation. The rank + nullity theorem

states that

dim ran (A) + dimker(A) = n.

To prove this, decompose

Rn = K ⊕K ′, K = kerA

Rk = R⊕R′, R = ranA.

Then

A : K ′ −→ R

is an isomorphism, and

n = dimK + dimK ′, k = dimR+ dimR′.

Then

A : K ′ −→ R

is an isomorphism, and

n = dimK + dimK ′, k = dimR+ dimR′.

Hence

n− k = (dimK + dimK ′)− (dimR+ dimR′)

= dimK − dimR′ = dimkerA− (dimRk − dim ranA).

Then

A : K ′ −→ R

is an isomorphism, and

n = dimK + dimK ′, k = dimR+ dimR′.

Hence

n− k = (dimK + dimK ′)− (dimR+ dimR′)

= dimK − dimR′ = dimkerA− (dimRk − dim ranA).

= ind(A),

where by definition, the index of A is defined to be

ind(A) = dimkerA− dimcokerA.

The first infinite dimensional example:

Let `2 denote the space of all square-summable sequences

(a0, a1, a2, . . .).

The first infinite dimensional example:

Let `2 denote the space of all square-summable sequences

(a0, a1, a2, . . .).

Define the left shift and right shift operators

L : `2 −→ `2, L((a0, a1, a2, . . .)) = (a1, a2, . . .),

and

R : `2 −→ `2, R((a0, a1, a2, . . .)) = (0, a0, a1, a2, . . .).

The first infinite dimensional example:

Let `2 denote the space of all square-summable sequences

(a0, a1, a2, . . .).

Define the left shift and right shift operators

L : `2 −→ `2, L((a0, a1, a2, . . .)) = (a1, a2, . . .),

and

R : `2 −→ `2, R((a0, a1, a2, . . .)) = (0, a0, a1, a2, . . .).

Then

dimkerL = 1, dimcokerL = 0

so

indL = 1.

On the other hand

dimkerR = 1, dimcokerR = 0

so

indR = −1.

On the other hand

dimkerR = 1, dimcokerR = 0

so

indR = −1.

Notice that

indLN = N, indRN = −N,

so we can construct operators with any given index.

Fredholm operators:

Let H1 and H2 be (separable) Hilbert spaces, and suppose

that A : H1 → H2 is a bounded operator.

Fredholm operators:

Let H1 and H2 be (separable) Hilbert spaces, and suppose

that A : H1 → H2 is a bounded operator.

Example:

A =1

i

d

acting on

H1 = H1(S1) = u :

∫|u′|2 dθ <∞, H2 = L2(S1),

A(einθ) = neinθ.

Definition:

A is Fredholm if kerA is finite dimensional, ranA is

closed and cokerA is finite dimensional.

If A is Fredholm, then

indA = dimkerA− dimcokerA ∈ Z

is well defined.

Definition:

A is Fredholm if kerA is finite dimensional, ranA is

closed and cokerA is finite dimensional.

If A is Fredholm, then

indA = dimkerA− dimcokerA ∈ Z

is well defined.

Intuition: If A is Fredholm, then

H1 = K ⊕K ′, H2 = R⊕R′,

where

A : K ′ −→ R

is an isomorphism and both K and R′ are finite dimensional.

A is Fredholm if it is ‘almost invertible’, or more precisely it

is invertible up to compact errors.

A is Fredholm if it is ‘almost invertible’, or more precisely it

is invertible up to compact errors.

Fredholm operators =

invertible elements in B(H1, H2)/K(H1, H2).

A is Fredholm if it is ‘almost invertible’, or more precisely it

is invertible up to compact errors.

Fredholm operators =

invertible elements in B(H1, H2)/K(H1, H2).

We can also make sense of what it means for an

unbounded, closed operator to be Fredholm.

A is Fredholm if it is ‘almost invertible’, or more precisely it

is invertible up to compact errors.

Fredholm operators =

invertible elements in B(H1, H2)/K(H1, H2).

We can also make sense of what it means for an

unbounded, closed operator to be Fredholm.

Recall that A : H1 → H2 is closed if

Graph(A) = (h,Ah) : h ∈ domain(A)

is a closed subspace of H1 ⊕H2.

A is Fredholm if it is ‘almost invertible’, or more precisely it

is invertible up to compact errors.

Fredholm operators =

invertible elements in B(H1, H2)/K(H1, H2).

We can also make sense of what it means for an

unbounded, closed operator to be Fredholm.

Recall that A : H1 → H2 is closed if

Graph(A) = (h,Ah) : h ∈ domain(A)

is a closed subspace of H1 ⊕H2.

(This is always the case if H1 = H2 = L2 and A is a

differential operator.)

Basic Fact:

If At is a one-parameter family of Fredholm operators

(depending continuously on t), then ind(At) is independent

of t.

Basic Fact:

If At is a one-parameter family of Fredholm operators

(depending continuously on t), then ind(At) is independent

of t.

Intuition:

At = tI −K, t > 0, K finite rank, symmetric

Then At gains a nullspace and a cokernel of the same

dimension every time t crosses an eigenvalue of K.

The Toeplitz index theorem:

Let H ⊂ L2(S1) be the Hilbert space consisting of all L2

functions on S1 which have only nonnegative Fourier

coefficients.

The Toeplitz index theorem:

Let H ⊂ L2(S1) be the Hilbert space consisting of all L2

functions on S1 which have only nonnegative Fourier

coefficients.

So

a ∈ H ⇔ a(θ) =

∞∑0

aneinθ,

∑|an|2 <∞.

The Toeplitz index theorem:

Let H ⊂ L2(S1) be the Hilbert space consisting of all L2

functions on S1 which have only nonnegative Fourier

coefficients.

So

a ∈ H ⇔ a(θ) =

∞∑0

aneinθ,

∑|an|2 <∞.

There is an orthogonal projection

Π : L2(S1) −→ H.

Suppose f(θ) is a continuous (complex-valued) function on

S1. Define the Toeplitz operator

Tf : H −→ H,

Tf(a) = ΠMfΠ,

where Mf is multiplication by f (bounded operator on

L2(S1)).

Suppose f(θ) is a continuous (complex-valued) function on

S1. Define the Toeplitz operator

Tf : H −→ H,

Tf(a) = ΠMfΠ,

where Mf is multiplication by f (bounded operator on

L2(S1)).

Question: When is Tf Fredholm? What is its index?

Suppose f(θ) is a continuous (complex-valued) function on

S1. Define the Toeplitz operator

Tf : H −→ H,

Tf(a) = ΠMfΠ,

where Mf is multiplication by f (bounded operator on

L2(S1)).

Question: When is Tf Fredholm? What is its index?

Theorem: The Toeplitz operator Tf is Fredholm if and only

if f is nowhere 0. In this case, its index is given by the

winding number of f (as a map S1 → C∗).

Part of proof: We guess that T1/f is a good approximation

to an inverse for Tf , so we compute:

TfT1/f = ΠMfΠM1/fΠ

Part of proof: We guess that T1/f is a good approximation

to an inverse for Tf , so we compute:

TfT1/f = ΠMfΠM1/fΠ

= ΠMfM1/fΠ + Π[Mf ,Π]M1/fΠ

Part of proof: We guess that T1/f is a good approximation

to an inverse for Tf , so we compute:

TfT1/f = ΠMfΠM1/fΠ

= ΠMfM1/fΠ + Π[Mf ,Π]M1/fΠ

= IH + Π[Mf ,Π]M1/fΠ.

The final term here is a compact operator! Obvious when f

has a finite Fourier series; for the general case, approximate

any continuous function by trigonometric polynomials.

Proof of Toeplitz index theorem:

Proof of Toeplitz index theorem:

First step: Let N be the winding number of f . Then there

is a continuous family ft, where each ft : S1 → C∗ is

continuous, such that f0 = f , f1 = eiNθ.

Proof of Toeplitz index theorem:

First step: Let N be the winding number of f . Then there

is a continuous family ft, where each ft : S1 → C∗ is

continuous, such that f0 = f , f1 = eiNθ.

ind(Tft) is independent of t, so it suffices to compute TeiNθ .

Proof of Toeplitz index theorem:

First step: Let N be the winding number of f . Then there

is a continuous family ft, where each ft : S1 → C∗ is

continuous, such that f0 = f , f1 = eiNθ.

ind(Tft) is independent of t, so it suffices to compute TeiNθ .

ΠMeiNθΠ :

∞∑n=0

aneinθ −→

∑∞

n=0 an−Neinθ if N < 0∑∞

n=N an−Neinθ if N ≥ 0

Proof of Toeplitz index theorem:

First step: Let N be the winding number of f . Then there

is a continuous family ft, where each ft : S1 → C∗ is

continuous, such that f0 = f , f1 = eiNθ.

ind(Tft) is independent of t, so it suffices to compute TeiNθ .

ΠMeiNθΠ :

∞∑n=0

aneinθ −→

∑∞

n=0 an−Neinθ if N < 0∑∞

n=N an−Neinθ if N ≥ 0

This corresponds to LN or RN , respectively. Thus

ind(Tf) = winding number of f

Elliptic operators

Let M be a smooth compact manifold of dimension n.

Elliptic operators

Let M be a smooth compact manifold of dimension n.

In any coordinate chart, a differential operator P on M has

the form

P =∑|α|≤m

aα(x)Dα,

where

α = (α1, . . . , αn), Dα =∂|α|

∂xα11 . . . ∂xαnn

.

Elliptic operators

Let M be a smooth compact manifold of dimension n.

In any coordinate chart, a differential operator P on M has

the form

P =∑|α|≤m

aα(x)Dα,

where

α = (α1, . . . , αn), Dα =∂|α|

∂xα11 . . . ∂xαnn

.

The symbol of P is the homogeneous polynomial

σm(P )(x; ξ) =∑|α|=m

aα(x)ξα, ξα = ξ

α11 . . . ξαnn .

If P is a system (i.e. acts between sections of two different

vector bundles E and F over M), then the coefficients

aα(x) lie in Hom(Ex, Fx).

If P is a system (i.e. acts between sections of two different

vector bundles E and F over M), then the coefficients

aα(x) lie in Hom(Ex, Fx).

P is called elliptic if

σm(P )(x; ξ) 6= 0 for ξ 6= 0,

(or is invertible as an element of Hom(Ex, Fx))

Fundamental Theorem: If P is an elliptic operator of order

m acting between sections of two vector bundles E and F

on a compact manifold M , then

P : L2(M ;E) −→ L2(M ;F )

defines a(n unbounded) Fredholm operator.

Fundamental Theorem: If P is an elliptic operator of order

m acting between sections of two vector bundles E and F

on a compact manifold M , then

P : L2(M ;E) −→ L2(M ;F )

defines a(n unbounded) Fredholm operator.

Hence ind(P ) is well defined.

Fundamental Theorem: If P is an elliptic operator of order

m acting between sections of two vector bundles E and F

on a compact manifold M , then

P : L2(M ;E) −→ L2(M ;F )

defines a(n unbounded) Fredholm operator.

Hence ind(P ) is well defined.

Question: Compute ind(P ).

Fundamental Theorem: If P is an elliptic operator of order

m acting between sections of two vector bundles E and F

on a compact manifold M , then

P : L2(M ;E) −→ L2(M ;F )

defines a(n unbounded) Fredholm operator.

Hence ind(P ) is well defined.

Question: Compute ind(P ).

One expects this to be possible because the index is

invariant under such a large class of deformations.

Interesting elliptic operators:

How about the Laplacian

∆0 : L2(M) −→ L2(M)

on functions?

Interesting elliptic operators:

How about the Laplacian

∆0 : L2(M) −→ L2(M)

on functions?

Nope: the Laplacian is self-adjoint,

〈∆0u, v〉 = 〈u,∆0v〉,

and so

ker∆0 = coker∆0 = R, hence ind(∆0) = 0.

The same is true for

∆k : L2Ωk(M) −→ L2Ωk(M)

∆k = dk−1d∗k + d∗k+1dk

since it too is self-adjoint.

The same is true for

∆k : L2Ωk(M) −→ L2Ωk(M)

∆k = dk−1d∗k + d∗k+1dk

since it too is self-adjoint.

In fact, the Hodge theorem states that

ker∆k = coker∆k

= Hk(M) (the space of harmonic forms),

and this is isomorphic to the singular cohomology,

Hksing(M,R).

Similarly

D = d+ d∗ : L2Ω∗(M) −→ L2Ω∗(M),

is still self-adjoint.

Similarly

D = d+ d∗ : L2Ω∗(M) −→ L2Ω∗(M),

is still self-adjoint.

However, we can break the symmetry:

Similarly

D = d+ d∗ : L2Ω∗(M) −→ L2Ω∗(M),

is still self-adjoint.

However, we can break the symmetry:

The Gauss-Bonnet operator:

DGB : L2Ωeven −→ L2Ωodd

Similarly

D = d+ d∗ : L2Ω∗(M) −→ L2Ω∗(M),

is still self-adjoint.

However, we can break the symmetry:

The Gauss-Bonnet operator:

DGB : L2Ωeven −→ L2Ωodd

kerDGB =⊕k

H2k(M),

cokerDGB =⊕k

H2k+1(M),

and so

indDGB =

n∑k=0

(−1)k dimHk(M) = χ(M),

the Euler characteristic of M .

and so

indDGB =

n∑k=0

(−1)k dimHk(M) = χ(M),

the Euler characteristic of M .

Note that ind(DGB) = 0 when dimM is odd.

and so

indDGB =

n∑k=0

(−1)k dimHk(M) = χ(M),

the Euler characteristic of M .

Note that ind(DGB) = 0 when dimM is odd.

This is not the index theorem, but only a Hodge-theoretic

calculation of the index of DGB in terms of something

familiar.

The Chern-Gauss-Bonnet theorem states that

χ(M) =

∫M

Pf(Ω).

The Chern-Gauss-Bonnet theorem states that

χ(M) =

∫M

Pf(Ω).

Here Pf(Ω) is an explicit differential form (called the

Pfaffian) which involves only the curvature tensor of M .

The Chern-Gauss-Bonnet theorem states that

χ(M) =

∫M

Pf(Ω).

Here Pf(Ω) is an explicit differential form (called the

Pfaffian) which involves only the curvature tensor of M .

One can think of this integral either as a differential

geometric quantity, or else as a characteristic number (the

evaluation of the Euler class of TM on the fundamental

homology class [M ]).

The signature operator:

Dsig = d+ d∗ : L2Ω+(M) −→ L2Ω−(M)

where

Ω±(M) is the ± eigenspace of an involution

τ : Ω∗(M) −→ Ω∗(M), τ2 = 1.

The signature operator:

Dsig = d+ d∗ : L2Ω+(M) −→ L2Ω−(M)

where

Ω±(M) is the ± eigenspace of an involution

τ : Ω∗(M) −→ Ω∗(M), τ2 = 1.

When dimM = 4k, then

indDsig = sign(H2k(M)×H2k(M)→ R),

which is a basic topological invariant, the signature of M .

The signature operator:

Dsig = d+ d∗ : L2Ω+(M) −→ L2Ω−(M)

where

Ω±(M) is the ± eigenspace of an involution

τ : Ω∗(M) −→ Ω∗(M), τ2 = 1.

When dimM = 4k, then

indDsig = sign(H2k(M)×H2k(M)→ R),

which is a basic topological invariant, the signature of M .

In all other cases, indDsig = 0.

The Hirzebruch signature theorem states that

sign(M) =

∫M

L(p),

where L(p) is the L-polynomial in the Pontrjagin classes.

This is another ‘curvature integral’, and also a

characteristic number. (It is a homotopy invariant of M .)

The (generalized) Dirac operator:

Let E± be two vector bundles over (M, g), and suppose that

there is a (fibrewise) multiplication

cl : TM × E± → E∓, Clifford multiplication

which satisfies

cl(v)cl(w) + cl(w)cl(v) = −2〈v, w〉Id.

Let E = E+ ⊕ E−.

The (generalized) Dirac operator:

Let E± be two vector bundles over (M, g), and suppose that

there is a (fibrewise) multiplication

cl : TM × E± → E∓, Clifford multiplication

which satisfies

cl(v)cl(w) + cl(w)cl(v) = −2〈v, w〉Id.

Let E = E+ ⊕ E−.

Let ∇ : C∞(M ;E)→ C∞(M ;E ⊗ T ∗M) be a connection.

Now define

D : L2(M ;E) −→ L2(M ;E)

by

D =

n∑i=1

cl(ei)∇ei ,

where e1, . . . , en is any local orthonormal basis of TM .

The Clifford relations show that

D2 = ∇∗∇+A

where A is a homomorphism.

The Clifford relations show that

D2 = ∇∗∇+A

where A is a homomorphism.

In other words,

“D2 is the Laplacian up to a term of order zero.”

In certain cases (when M is a spin manifold: w2 = 0), there

is a distinguished Dirac operator D acting between sections

of the spinor bundle S over M .

One can then ‘twist’ this basic Dirac operator

D : L2(M ;S) −→ L2(M ;S)

by tensoring with an arbitrary vector bundle E, to get the

twisted Dirac operator

DE = D ⊗ 1⊕ cl(1⊗∇E) : L2(M : S ⊗ E) −→ L2(M ;S ⊗ E).

The index theorem for twisted Dirac operators states that

ind(DE) =

∫M

A(TM)ch(E).

The integrand is again a ‘curvature integral’; it is the

product of the A genus of M and the Chern character of

the bundle E.

The index theorem for twisted Dirac operators states that

ind(DE) =

∫M

A(TM)ch(E).

The integrand is again a ‘curvature integral’; it is the

product of the A genus of M and the Chern character of

the bundle E.

Both DGB and Dsig are twisted Dirac operators, and both

Pf(R) and L(p) are the product of A(TM) and ch(E) for

appropriate bundles E.

There is a general index theorem for arbitrary elliptic

operators P (which are not necessarily of order 1), acting

between sections of two bundles E and F . This takes the

form

ind(P ) =

∫M

R(P,E, F ).

The right hand side here is a characteristic number for

some bundle associated to E, F and the symbol of P .

By Chern-Weil theory, this can be written as a curvature

integral, but this is less natural now.

Intermission:

One day at the watering hole, an elephant looked around

and carefully surveyed the turtles in view. After a few

seconds thought, he walked over to one turtle, raised his

foot, and KICKED the turtle as far as he could. (Nearly a

mile) A watching hyena asked the elephant why he did it?

”Well, about 30 years ago I was walking through a stream

and a turtle bit my foot. Finally I found the S.O.B and

repaid him for what he had done to me.” ”30 years!!! And

you remembered...But HOW???”

Intermission:

One day at the watering hole, an elephant looked around

and carefully surveyed the turtles in view. After a few

seconds thought, he walked over to one turtle, raised his

foot, and KICKED the turtle as far as he could. (Nearly a

mile) A watching hyena asked the elephant why he did it?

”Well, about 30 years ago I was walking through a stream

and a turtle bit my foot. Finally I found the S.O.B and

repaid him for what he had done to me.” ”30 years!!! And

you remembered...But HOW???”

”I have turtle recall.”

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

2. It has good ‘multiplicative’ and excision properties

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

2. It has good ‘multiplicative’ and excision properties

3. It defines an additive homomorphism on K0(M).

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

2. It has good ‘multiplicative’ and excision properties

3. It defines an additive homomorphism on K0(M).

Here we regard E → indDE as a semigroup homomorphism

vector bundles on M −→ Z

and then extend to the formal group, which is K0(M).

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

2. It has good ‘multiplicative’ and excision properties

3. It defines an additive homomorphism on K0(M).

Here we regard E → indDE as a semigroup homomorphism

vector bundles on M −→ Z

and then extend to the formal group, which is K0(M).

4. It is invariant, in a suitable sense, under cobordisms of M

The original proofs

The index is computable because it is invariant under many

big alterations and behaves well with respect to many

natural operations:

1. It is invariant under deformations of P amongst

Fredholm operators

2. It has good ‘multiplicative’ and excision properties

3. It defines an additive homomorphism on K0(M).

Here we regard E → indDE as a semigroup homomorphism

vector bundles on M −→ Z

and then extend to the formal group, which is K0(M).

4. It is invariant, in a suitable sense, under cobordisms of M

Hirzebruch (1957) proved the signature theorem for smooth

algebraic varieties, using the cobordism invariance of the

signature and the L-class.

Hirzebruch (1957) proved the signature theorem for smooth

algebraic varieties, using the cobordism invariance of the

signature and the L-class.

An elaboration of this same strategy may be used to reduce

the computation of the index of a general elliptic operator

P to that of a twisted Dirac operator on certain very

special manifolds (products of complex projective spaces).

Steps of the proof

Steps of the proof

1. ‘Deform’ the mth order elliptic operator P to a 1st order

twisted Dirac operator D.

Steps of the proof

1. ‘Deform’ the mth order elliptic operator P to a 1st order

twisted Dirac operator D.

This requires the use of pseudodifferential operators

(polynomials are too rigid), and was one of the big early

motivations for the theory of pseudodifferential operators.

It also requires K-theory.

Steps of the proof

1. ‘Deform’ the mth order elliptic operator P to a 1st order

twisted Dirac operator D.

This requires the use of pseudodifferential operators

(polynomials are too rigid), and was one of the big early

motivations for the theory of pseudodifferential operators.

It also requires K-theory.

2. Replace the twisted Dirac operator on the manifold M

by a twisted Dirac operator on a new manifold M ′ such that

M −M ′ = ∂W . The index stays the same!

3. Check that the two sides are equal on ‘enough’

examples, products of complex projective spaces (these

generate the cobordism ring). This requires the

multiplicativity of the index.

A simplification:

Instead of the final two steps, it is in many ways preferable

to compute the index of twisted Dirac operators on an

arbitrary manifold M directly.

A simplification:

Instead of the final two steps, it is in many ways preferable

to compute the index of twisted Dirac operators on an

arbitrary manifold M directly.

Later proofs involve verifying that the two sides of the

index theorem on the same directly for Dirac operators on

arbitrary manifolds; this still uses K-theory and

pseudodifferential operators, but obviates the bordism.

A simplification:

Instead of the final two steps, it is in many ways preferable

to compute the index of twisted Dirac operators on an

arbitrary manifold M directly.

Later proofs involve verifying that the two sides of the

index theorem on the same directly for Dirac operators on

arbitrary manifolds; this still uses K-theory and

pseudodifferential operators, but obviates the bordism.

From now on I’ll focus only on the index theorem for

twisted Dirac operators.

The heat equation approach

Suppose that P is a nonnegative second order self-adjoint

elliptic operator.

The heat equation approach

Suppose that P is a nonnegative second order self-adjoint

elliptic operator.

The examples I have in mind are of the form

P = D∗D or P = DD∗

The heat equation approach

Suppose that P is a nonnegative second order self-adjoint

elliptic operator.

The examples I have in mind are of the form

P = D∗D or P = DD∗

Note that if D = d+ d∗ on Ωeven or Ω+, then D = d+ d∗ on

Ωodd or Ω−, so

D∗D = (d+ d∗)2 = dd∗ + d∗d, DD∗ = dd∗ + d∗d

on Ωeven (Ω+), and Ωodd (Ω−), respectively.

The heat equation approach

Suppose that P is a nonnegative second order self-adjoint

elliptic operator.

The examples I have in mind are of the form

P = D∗D or P = DD∗

Note that if D = d+ d∗ on Ωeven or Ω+, then D = d+ d∗ on

Ωodd or Ω−, so

D∗D = (d+ d∗)2 = dd∗ + d∗d, DD∗ = dd∗ + d∗d

on Ωeven (Ω+), and Ωodd (Ω−), respectively.

The initial value problem for the heat equation is(∂

∂t+ P

)u = 0 on M × [0,∞),

with initial data u(x,0) = u0(x).

The initial value problem for the heat equation is(∂

∂t+ P

)u = 0 on M × [0,∞),

with initial data u(x,0) = u0(x).

The solution is given by the heat kernel H

u(x, t) = (Hu0)(x, t) =

∫M

H(x, y, t)φ(y) dVy.

Let φj , λj be the eigendata for P , i.e.

Pφj = λjφj , φj dense in L2(M).

Let φj , λj be the eigendata for P , i.e.

Pφj = λjφj , φj dense in L2(M).

Then

H(x, y, t) =

∞∑j=0

e−λjtφj(x)φj(y)

because

u0 =

∞∑j=0

cjφj(x)⇒∫M

∞∑j=0

e−λjtφj(x)φj(y)

∞∑j′=0

cj′φj′(y) dVy

Let φj , λj be the eigendata for P , i.e.

Pφj = λjφj , φj dense in L2(M).

Then

H(x, y, t) =

∞∑j=0

e−λjtφj(x)φj(y)

because

u0 =

∞∑j=0

cjφj(x)⇒∫M

∞∑j=0

e−λjtφj(x)φj(y)

∞∑j′=0

cj′φj′(y) dVy

=

∞∑j=0

cje−λjtφj(x).

The McKean-Singer argument

Let (φj , λj) be the eigendata for D∗D and (ψ`, µ`) be the

eigendata for DD∗.

The McKean-Singer argument

Let (φj , λj) be the eigendata for D∗D and (ψ`, µ`) be the

eigendata for DD∗.

Spectral cancellation

Let E(λ) = φ : D∗Dφ = λφ and F (µ) = ψ : DD∗φ = µφ.Then

for λ 6= 0, D : E(λ) −→ F (λ)

is an isomorphism.

The McKean-Singer argument

Let (φj , λj) be the eigendata for D∗D and (ψ`, µ`) be the

eigendata for DD∗.

Spectral cancellation

Let E(λ) = φ : D∗Dφ = λφ and F (µ) = ψ : DD∗φ = µφ.Then

for λ 6= 0, D : E(λ) −→ F (λ)

is an isomorphism.

Proof:

D∗Dφ = λφ⇒ D(D∗Dφ) = (DD∗)(Dφ) = λ(Dφ)

so λ is an eigenvalue for DD∗.

Also need to check that Dφ 6= 0:

Dφ = 0⇒ 0 = ||Dφ||2 = 〈D∗Dφ, φ〉 ⇒ φ ∈ kerD.

Also need to check that Dφ 6= 0:

Dφ = 0⇒ 0 = ||Dφ||2 = 〈D∗Dφ, φ〉 ⇒ φ ∈ kerD.

Conclusion The heat kernel traces are given by the sums

Tr(e−tD∗D) =

∞∑j=0

e−λjt

Tr(e−tDD∗) =

∞∑`=0

e−µ`t.

Also need to check that Dφ 6= 0:

Dφ = 0⇒ 0 = ||Dφ||2 = 〈D∗Dφ, φ〉 ⇒ φ ∈ kerD.

Conclusion The heat kernel traces are given by the sums

Tr(e−tD∗D) =

∞∑j=0

e−λjt

Tr(e−tDD∗) =

∞∑`=0

e−µ`t.

Tr(e−tD∗D)−Tr(e−tDD

∗) =

∑λj=0

e−λjt −∑µ`=0

e−µ`t

= dimkerD∗D − dimkerDD∗ = ind(D)!

Also need to check that Dφ 6= 0:

Dφ = 0⇒ 0 = ||Dφ||2 = 〈D∗Dφ, φ〉 ⇒ φ ∈ kerD.

Conclusion The heat kernel traces are given by the sums

Tr(e−tD∗D) =

∞∑j=0

e−λjt

Tr(e−tDD∗) =

∞∑`=0

e−µ`t.

Tr(e−tD∗D)−Tr(e−tDD

∗) =

∑λj=0

e−λjt −∑µ`=0

e−µ`t

= dimkerD∗D − dimkerDD∗ = ind(D)!

In particular,

Tr(e−tD∗D)−Tr(e−tDD

∗) is independent of t!

Localization as t 0

The heat kernel H(x, y, t) measures ‘the amount of heat at

x ∈M at time t, if an initial unit (delta function) of heat is

placed at y ∈M at time 0.

Localization as t 0

The heat kernel H(x, y, t) measures ‘the amount of heat at

x ∈M at time t, if an initial unit (delta function) of heat is

placed at y ∈M at time 0.

As t 0, if x 6= y, then H(x, y, t)→ 0 (exponentially

quickly).

Localization as t 0

The heat kernel H(x, y, t) measures ‘the amount of heat at

x ∈M at time t, if an initial unit (delta function) of heat is

placed at y ∈M at time 0.

As t 0, if x 6= y, then H(x, y, t)→ 0 (exponentially

quickly).

On the other hand, along the diagonal

H(x, x, t) ∼∞∑j=0

ajt−n/2+j .

The coefficients aj are called the heat traceinvariants.

They are local, i.e. they can be written as

aj =

∫M

pj(x) dVx,

where pj is a universal polynomial in the coefficients of

the operator.

They are local, i.e. they can be written as

aj =

∫M

pj(x) dVx,

where pj is a universal polynomial in the coefficients of

the operator.

If P is a geometric differential operator, then the pj are

universal functions of the metric and its covariant

derivatives up to some order.

ind(D) = Tre−tD∗D −Tre−tDD

=

∞∑j=0

aj(D∗D)t−n/2+j −

∞∑j=0

aj(DD∗)t−n/2+j

= an/2(D∗D)− an/2(DD∗).

ind(D) = Tre−tD∗D −Tre−tDD

=

∞∑j=0

aj(D∗D)t−n/2+j −

∞∑j=0

aj(DD∗)t−n/2+j

= an/2(D∗D)− an/2(DD∗).

Note: this implies that the index is zero when n/2 /∈N!

ind(D) = Tre−tD∗D −Tre−tDD

=

∞∑j=0

aj(D∗D)t−n/2+j −

∞∑j=0

aj(DD∗)t−n/2+j

= an/2(D∗D)− an/2(DD∗).

Note: this implies that the index is zero when n/2 /∈N!

The difficulty:

The term an/2 is high up in the asymptotic expansion, and

so very hard to compute explicitly!

V.K. Patodi did the computation explicitly for a few basic

examples (Gauss-Bonnet, etc.). He discovered a miraculous

cancellation which made the computation tractable.

His computation was extended somewhat by

Atiyah-Bott-Patodi (1973)

P. Gilkey gave another proof based on geometric invariant

theory, i.e. the idea that these universal polynomials are

SO(n)-invariant, in an appropriate sense, and hence must

be polynomials in the Pontrjagin forms. Then, you calculate

the coefficients by checking on lots of examples ....

The big breakthrough was accomplished by Getzler (as an

undergraduate!):

The big breakthrough was accomplished by Getzler (as an

undergraduate!):

he defined a rescaling of the Clifford algebra and the Clifford

bundles in terms of the variable t, which as the effect that

the ‘index term’ an/2 occurs as the leading coefficient!

The first application: (non)existence of metrics of positive

scalar curvature

Suppose (M2n, g) is a spin manifold. Thus the Dirac

operator D acts between sections of the spinor bundles S±.

D : L2(M ;S+) −→ L2(M ;S−).

The first application: (non)existence of metrics of positive

scalar curvature

Suppose (M2n, g) is a spin manifold. Thus the Dirac

operator D acts between sections of the spinor bundles S±.

D : L2(M ;S+) −→ L2(M ;S−).

The Weitzenbock formula states that

D∗D = DD∗ = ∇∗∇+R

4

The first application: (non)existence of metrics of positive

scalar curvature

Suppose (M2n, g) is a spin manifold. Thus the Dirac

operator D acts between sections of the spinor bundles S±.

D : L2(M ;S+) −→ L2(M ;S−).

The Weitzenbock formula states that

D∗D = DD∗ = ∇∗∇+R

4

Corollary: (Lichnerowicz) If M admits a metric g of positive

scalar curvature, then kerD = cokerD = 0. Hence indD = 0,

and so the A-genus of M vanishes.

The Gromov-Lawson conjecture states that a slight

generalization of this condition (the vanishing of the

corresponding integral characteristic class, and suitably

modified to allow odd dimensions) is necessary and

sufficient for the existence of a metric of positive scalar

curvature (at least when π1(M) = 0).

The Gromov-Lawson conjecture states that a slight

generalization of this condition (the vanishing of the

corresponding integral characteristic class, and suitably

modified to allow odd dimensions) is necessary and

sufficient for the existence of a metric of positive scalar

curvature (at least when π1(M) = 0).

Proved (when M simply connected) if the dimension is ≥ 5

(Stolz).

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

2. Pseudoholomorphic curves

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

2. Pseudoholomorphic curves

3. Complex structures

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

2. Pseudoholomorphic curves

3. Complex structures

Deformation theory: given a solution u0 to this equation,

find all ‘nearby’ solutions u.

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

2. Pseudoholomorphic curves

3. Complex structures

Deformation theory: given a solution u0 to this equation,

find all ‘nearby’ solutions u.

Write

N(u) = DN |u0 (u− u0) +Q(u− u0).

Another application: computing dimensions of moduli

spaces:

Suppose N is a nonlinear elliptic operator such that the

solutions of N(u) = 0 correspond to some interesting

geometric objects, e.g.

1. Self-dual Yang-Mills connections

2. Pseudoholomorphic curves

3. Complex structures

Deformation theory: given a solution u0 to this equation,

find all ‘nearby’ solutions u.

Write

N(u) = DN |u0 (u− u0) +Q(u− u0).

If the linearization L = DN |u0, the implicit function theorem

implies that

if L is surjective

then, in a neighbourhood of u0,

solutions of N(u) = 0↔ solutions of Lφ = 0 = kerL.

If the linearization L = DN |u0, the implicit function theorem

implies that

if L is surjective

then, in a neighbourhood of u0,

solutions of N(u) = 0↔ solutions of Lφ = 0 = kerL.

Best situation (unobstructed deformation theory) is when L

surjective, so that

indL = dimkerL.

This is computable, and is the local dimension of solution

space.

Example:

Let M4 be compact and E a rank 2 bundle over M with

charge c2(E) = k ∈ Z. Then the moduli space Mk of all

self-dual Yang-Mills instantons of charge k ‘should be’ (and

is, for generic choice of background metric) a smooth

manifold, with

dimMk = 8c2(E)− 3(1− β1(M) + β+2 (M)).

Example:

Let M4 be compact and E a rank 2 bundle over M with

charge c2(E) = k ∈ Z. Then the moduli space Mk of all

self-dual Yang-Mills instantons of charge k ‘should be’ (and

is, for generic choice of background metric) a smooth

manifold, with

dimMk = 8c2(E)− 3(1− β1(M) + β+2 (M)).

Freedom to choose generic metric means one can always

assume one is in the unobstructed situation.

The index formula for manifolds with boundary

The index formula for manifolds with boundary

Early version (Atiyah-Bott) applies to elliptic operators with

local (e.g. Dirichlet or Neumann) boundary conditions.

The index formula for manifolds with boundary

Early version (Atiyah-Bott) applies to elliptic operators with

local (e.g. Dirichlet or Neumann) boundary conditions.

Not available for Dirac-type operators

The index formula for manifolds with boundary

Early version (Atiyah-Bott) applies to elliptic operators with

local (e.g. Dirichlet or Neumann) boundary conditions.

Not available for Dirac-type operators

Atiyah-Patodi-Singer (1975):

Suppose that Y = ∂M has a collar neighbourhood

[0,1)× Y ⊂M , so that the metric, the bundles, the

connection, are all of product form in this neighbourhood.

The index formula for manifolds with boundary

Early version (Atiyah-Bott) applies to elliptic operators with

local (e.g. Dirichlet or Neumann) boundary conditions.

Not available for Dirac-type operators

Atiyah-Patodi-Singer (1975):

Suppose that Y = ∂M has a collar neighbourhood

[0,1)× Y ⊂M , so that the metric, the bundles, the

connection, are all of product form in this neighbourhood.

Then

D = cl(en)

(∂

∂xn+A

),

where

A : L2(Y ;S) −→ L2(Y ;S)

is self-adjoint!

The index formula for manifolds with boundary

Early version (Atiyah-Bott) applies to elliptic operators with

local (e.g. Dirichlet or Neumann) boundary conditions.

Not available for Dirac-type operators

Atiyah-Patodi-Singer (1975):

Suppose that Y = ∂M has a collar neighbourhood

[0,1)× Y ⊂M , so that the metric, the bundles, the

connection, are all of product form in this neighbourhood.

Then

D = cl(en)

(∂

∂xn+A

),

where

A : L2(Y ;S) −→ L2(Y ;S)

is self-adjoint!

Let φj , λj be the eigendata for A,

Aφj = λjφj .

λj → ±∞.

Let φj , λj be the eigendata for A,

Aφj = λjφj .

λj → ±∞.

Decompose

L2(Y ;S) = H+ ⊕H−,

where H+ is the sum of the eigenspaces corresponding to

positive eigenvalues and H− is the sum of the eigenspaces

corresponding to negative eigenvalues.

Let φj , λj be the eigendata for A,

Aφj = λjφj .

λj → ±∞.

Decompose

L2(Y ;S) = H+ ⊕H−,

where H+ is the sum of the eigenspaces corresponding to

positive eigenvalues and H− is the sum of the eigenspaces

corresponding to negative eigenvalues.

Π : L2(Y ;S) −→ H+

is the orthogonal projection.

Let φj , λj be the eigendata for A,

Aφj = λjφj .

λj → ±∞.

Decompose

L2(Y ;S) = H+ ⊕H−,

where H+ is the sum of the eigenspaces corresponding to

positive eigenvalues and H− is the sum of the eigenspaces

corresponding to negative eigenvalues.

Π : L2(Y ;S) −→ H+

is the orthogonal projection.

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

This is Fredholm

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

This is Fredholm

The nonlocal boundary condition given by Π is known as

the APS boundary condition.

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

This is Fredholm

The nonlocal boundary condition given by Π is known as

the APS boundary condition.

To state the index formula, we need one more is definition:

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

This is Fredholm

The nonlocal boundary condition given by Π is known as

the APS boundary condition.

To state the index formula, we need one more is definition:

η(s) =∑λj 6=0

(sgnλj)|λj |−s.

The Atiyah-Patodi-Singer boundary problem:

(D,Π) : L2(X,S+) −→ L2(X;S−)⊕ L2(Y ;S)

(D,Π)u = (Du,Π u|Y ).

This is Fredholm

The nonlocal boundary condition given by Π is known as

the APS boundary condition.

To state the index formula, we need one more is definition:

η(s) =∑λj 6=0

(sgnλj)|λj |−s.

Weyl eigenvalue asymptotics for D implies that η(s) is

holomorphic for Re(s) > n.

Theorem: The eta function extends meromorphically to the

whole complex plane. It is regular at s = 0.

Theorem: The eta function extends meromorphically to the

whole complex plane. It is regular at s = 0.

Definition: The eta invariant of A, η(A) is by definition the

value of this eta function at s = 0.

Theorem: The eta function extends meromorphically to the

whole complex plane. It is regular at s = 0.

Definition: The eta invariant of A, η(A) is by definition the

value of this eta function at s = 0.

The APS index theorem:

ind(D,Π) =

∫M

A(TM)ch(E)−1

2(η(A) + h),

where h = dimkerA.

Theorem: The eta function extends meromorphically to the

whole complex plane. It is regular at s = 0.

Definition: The eta invariant of A, η(A) is by definition the

value of this eta function at s = 0.

The APS index theorem:

ind(D,Π) =

∫M

A(TM)ch(E)−1

2(η(A) + h),

where h = dimkerA.

This suggests that η(A) is in many interesting ways the

odd-dimensional generalization of the index.

Theorem: The eta function extends meromorphically to the

whole complex plane. It is regular at s = 0.

Definition: The eta invariant of A, η(A) is by definition the

value of this eta function at s = 0.

The APS index theorem:

ind(D,Π) =

∫M

A(TM)ch(E)−1

2(η(A) + h),

where h = dimkerA.

This suggests that η(A) is in many interesting ways the

odd-dimensional generalization of the index.

The APS index formula is then a sort of transgression

formula, connecting the index in even dimensions and the

eta invariant in odd dimensions.

top related