dirac operators on manifolds and applications to physics · 1 introduction this report develops the...

33
Dirac Operators on Manifolds and Applications to Physics Angus Alexander Supervised by Adam Rennie and Alan Carey University of Wollongong Vacation Research Scholarships are funded jointly by the Department of Education and Training and the Australian Mathematical Sciences Institute.

Upload: others

Post on 25-May-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Dirac Operators on Manifolds and

Applications to Physics

Angus AlexanderSupervised by Adam Rennie and Alan Carey

University of Wollongong

Vacation Research Scholarships are funded jointly by the Department of Education and Training

and the Australian Mathematical Sciences Institute.

Page 2: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Abstract

This work was completed in the summer of 2017-18 as part of an AMSI vacation research scholarship

under the supervision of Adam Rennie and Alan Carey. The aim of the project was to study differential

operators on manifolds and applications to physics, in particular to understand the APS boundary conditions

presented by Bar and Strohmaier in [2] and apply them to the case of a cylinder. We further attempt to

combine the work in [2] with that of Connes’ in [3] and develop some possible candidates for a trace in

relativistic quantum mechanics.

1 Introduction

This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and applies

the theory to the case of a cylinder with APS boundary conditions as found in [2].

In Section 2 we develop the theory of elliptic differential operators on manifolds using unbounded operator

theory for Hilbert spaces. In particular, we extend the notion of an adjoint to a differential operator and prove

some results relating to boundary conditions and self-adjointness.

In Section 3 we develop the algebraic prerequisites required for understanding Dirac operators. The section

is focused on Clifford algebras, modules and bundles with a number of examples. The concepts of spinor bundles

and connections are also introduced.

In Section 4, we define a Dirac operator for a vector bundle over a manifold and provide numerous Euclidean

examples, as well as the Dirac operator on a Riemannian cylinder.

In Section 5, the theory developed in the previous sections is utilised to determine the Dirac operator on a

Lorentzian cylinder. The eigenvalues for the operators are computed, as well as some candidates for a trace in

an attempt to combine the work in [2] and that of Connes’ in [3].

2 Elliptic Differential Operators

The material in this section is mostly based on [5, Chapter 10], with some elements drawn from [10] and [7]. We

develop the theory of differential operators on smooth vector bundles over smooth manifolds. For a refresher

on vector bundles and manifolds, see [4] and [6].

2.1 First Order Differential Operators

Definition 2.1. Let M be a smooth manifold and let p ∶ E → M be a smooth vector bundle over M . Let

C∞(M,E) be the space of smooth sections of E. A first order linear differential operator on E is a complex

linear map D ∶ C∞(M,E)→ C∞(M,E) which satisfies the following two properties:

1. If u1 and u2 are smooth sections of E which agree on an open set U ⊆M then Du1 and Du2 agree on U ,

and

1

Page 3: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

2. For each coordinate patch U ⊆M , if we choose coordinates xj in U and a trivialisation for the bundle E

over U , then D can be represented in local coordinates by

Du =∑j

Aj∂u

∂xj+Bu,

where the Aj and B are smooth matrix valued functions on U .

We define the support of D to be the complement of the largest open set U for which the restriction of D to U

is zero.

The functions Aj and B above are dependent on the bundle trivialisation, however retain their form. Let

ΦU ∶ E∣U → U × Ck and ΦV ∶ E∣V → V × Ck be bundle trivialisations. The transition function gUV ∶ U ∩ V →

GLk(C) is defined as

gUV = ΦU ○Φ−1V .

Given coordinates ϕU ∶ U → Rn and u ∈ C∞(M,E∣U), we have for x ∈ Rn that

ΦU(Du)(x) =∑j

Aj(x)∂

∂xjΦU(u) ○ ϕ−1U (x) +BΦu(u) ○ ϕ−1U (x).

Let y = ϕV ○ ϕ−1U (x) and compute that

gV U (ΦU(Du)) (ϕ−1U (x)) = ΦV (Du) (ϕ−1V (y))

=∑j

(gV UAj∂

∂xj(ΦU ○Φ−1

V ○ΦV u)) (ϕ−1V ○ ϕV ○ ϕ−1U (x))

+ (gV UBΦU ○Φ−1V ○ΦV u) (ϕ−1v ○ ϕV ○ ϕ−1U (x))

=∑j

(gV UAj∂gUV∂xj

(ϕ−1U (x)) (ΦV (u)) (ϕ−1V (y)))

+∑j

gV UAjgUV∂yk

∂xj∂ΦV (u)∂yk

○ ϕ−1V (y) + (gV UBgUV ΦV (u)) (ϕ−1V (y))

=∑j

Aj∂ΦV (u)∂yj

(ϕ−1V (y)) + Bu (ϕ−1V (y)) .

From this expression we see that in the new trivialisation we still have a differential operator, however the

functions Aj and B have changed.

Definition 2.2. Define the symbol of D, σD ∶ T ∗M → End(E) by

σD(x, ξ) = i∑j

Ajξj .

Remark 2.3. Definition 2.1 only refers to first order differential operators. We may more generally consider

an order k differential operator by replacing condition (2) with

Du = ∑∣α∣≤k

Aα∂ ∣α∣u

∂xα,

2

Page 4: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

where α = (α1, . . . , αn) ∈ Nn is a multi-index and ∣α∣ =n

∑j=1

αj . We may also extend Definition 2.2 to an order k

operator as

σD(x, ξ) = ik ∑∣α∣=k

Aαξα.

Definition 2.1 can also be extended to differential operators between different vector bundles E and F as

D ∶ C∞(M,E) → C∞(M,F ), where in local coordinates D is given by the same formula as above. The symbol

is defined in this case in the same manner, however σD ∶ T ∗M → Hom(E,F ). Note also that from the definition,

it is clear that if D1 and D2 are differential operators, then

σD1○D2(x, ξ) = σD1(x, ξ) ○ σD2(x, ξ).

A number of differential operators are already familiar from vector calculus and Riemannian geometry.

Example 2.4. Recall that the Laplace-Beltrami operator ∆ ∶ C∞(M,E)→ C∞(M,E) is given in local coordi-

nates as (for f ∈ C∞(M,E))

∆f = 1√g∑j,k

∂xj(√ggjk ∂f

∂xk)

=∑j,k

(gjk ∂2f

∂xj∂xk+ 1

√g

∂√g

∂xj

∂f

∂xk).

So we calculate that for (x, ξ) ∈ T ∗M ,

σD(x, ξ) = i2∑j,k

gjkξjξk

= − ∥ξ∥2 .

Definition 2.5. If the vector bundle E has a Hermitian metric (⋅, ⋅) and M has a smooth measure µ, then we

may define an inner product on C∞c (M,S) by

⟨u, v⟩ = ∫M

(u(x), v(x))dµ(x).

Using this we define L2(M,S) to be the completion of C∞c (M,S) relative to this inner product.

Proposition 2.6. Let D ∶ C∞(M,E) → C∞(M,E) be a differential operator. There is a unique differential

operator D� ∶ C∞(M,E)→ C∞(M,E) such that

⟨Du, v⟩ = ⟨u,D�v⟩

for all u, v ∈ C∞c (M,E). The symbols of D and D� are related by the formula

σD�(x, ξ) = σD(x, ξ)∗.

We call D� the adjoint of D.

3

Page 5: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Proof. By using a partition of unity, it suffices to show existence in the case where D is supported in a single

coordinate patch. Then we compute that on this patch U we have

⟨Du, v⟩ = ∫U

⎛⎝∑j

Aj∂u

∂xj+Bu, v

⎞⎠

dx

=∑j∫U(Aj

∂u

∂xj, v)dx + ∫

U(Bu, v)dx

=∑j∫U( ∂u∂xj

,A∗j v)dx + ∫

U(u,B∗v)dx

=∑j∫U[ ∂

∂xj(u,A∗

j v) − (u, ∂

∂xj[A∗

j v])]dx + ∫U(u,B∗v)dx

= −∑j∫U(u,

∂A∗j

∂xjv +A∗

j

∂v

∂xj)dx + ∫

U(u,B∗v)dx

= ∫U

⎛⎝u,∑

j

−A∗j

∂v

∂xj+⎡⎢⎢⎢⎣B∗ −∑

j

∂Aj

∂xj

⎤⎥⎥⎥⎦v⎞⎠

dx,

where the second last equality follows from the fact that D is zero in a neighbourhood of the boundary of U .

For uniqueness, suppose that D1 and D2 are both formal adjoints for D. Define L =D1 −D2 and compute that

for any u, v ∈ C∞c (M,E) we have

⟨u,Lv⟩ = ⟨u, (D1 −D2)v⟩

= ⟨u,D1v⟩ − ⟨u,D2v⟩

= ⟨Du, v⟩ − ⟨Du, v⟩

= 0.

Since the range of L is orthogonal to every u ∈ C∞c (M,E), we have L = 0 as required. Note also that from the

above formula for D�, we see that for (x, ξ) ∈ T ∗M we have

σD�(x, ξ) = i∑j

−A∗j ξj

=⎛⎝i∑j

Ajξj⎞⎠

= σD(x, ξ)∗.

Example 2.7. Let Mg denote the operator of multiplication by g in C∞(M,S). Then we calculate that for

4

Page 6: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

any smooth section u of E

([D,Mg])u =D(gu) − gDu

=∑j

(Aj∂(gu)∂xj

) +Bgu − g⎛⎝∑j

Aj∂u

∂xj+Bu

⎞⎠

=∑j

(Aj∂g

∂xju +Ajg

∂u

∂xj− gAj

∂u

∂xj) +Bgu − gBu

=⎛⎝∑j

Aj∂g

∂xj

⎞⎠u.

Evaluating this computation at x ∈M , we have

i[D,Mg]u(x) = σD(x,dg)u(x).

Henceforth we shall refer to a first order differential operator as simply a differential operator.

2.2 Symmetric and Self-Adjoint Differential Operators

In this section we develop the Hilbert space theory of differential operators, although much care is needed since

these operators are usually not bounded. We consider a differential operator D as an unbounded operator

initially defined on C∞c (M,E) ⊂ L2(M,E). We begin with some technical results.

Lemma 2.8. Let D ∶ C∞c (M,E)→ C∞

c (M,E) be a differential operator. Then D is closable.

Proof. Suppose that vj → 0 in C∞c (M,E) and Dvj → w ∈ L2(M,E). For any u ∈ C∞

c (M,E) we have that

⟨u,w⟩ = limj→∞

⟨u,Dvj⟩

= limj→∞

⟨D�u, vj⟩

= ⟨D�u,0⟩

= 0.

Since w is orthogonal to C∞c (M,E), a dense subspace of L2(M,E), we have that w = 0 and so D is closable.

The overline indicating closure of an operator will be omitted when the distinction between the two is not

necessary. Recall that an operator T is symmetric if ⟨Tu, v⟩ = ⟨u,Tv⟩ for all u, v ∈ Dom(T ). For a differential

operator on a manifold without boundary this amounts to being self-adjoint, and these are the primary operators

of interest to us.

Definition 2.9. Let D ∶ C∞c (M,E) → C∞

c (M,E) be a symmetric differential operator. The minimal domain

of D is Dom(D) and the maximal domain of D is Dom(D∗).

Note that if D is symmetric then D ⊆ D ⊆ D∗ and all three of these operators may be different from one

another.

5

Page 7: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Example 2.10. Consider the manifold M = (0,1). Let Dom(D) = C∞c (M,C) and define the differential

operator D ∶ Dom(D)→ L2([0,1]) by D = i d

dx. Recall that

Dom(D∗) = {ξ ∈ L2([0,1]) ∶ there exists η ∈ L2([0,1]) s.t for all ψ ∈ Dom(D), ⟨Dψ, ξ⟩ = ⟨ψ, η⟩} .

Then we may check that ξ(x) = x is in Dom(D∗), since for any ψ ∈ Dom(D),

⟨Dψ, ξ⟩ = i∫1

0

dxxdx

= i∫1

0

d

dx[ψ(x)x]dx − i∫

1

0ψ(x)dx

= ⟨ψ,−i⟩.

However ξ ∉ Dom(D). To see this, first fix u ∈ Dom(D) and calculate that

⟨Du,1⟩ = i∫1

0

du

dxdx

= 0,

then extend by continuity in the graph norm (∥ξ∥Gr = ∥ξ∥ + ∥Dξ∥) to see that this holds for all u ∈ Dom(D)

also. Now note that

⟨Dξ,1⟩ = i∫1

01 dx

= i

≠ 0.

Hence D∗ and D are distinct operators.

The example above suggests that failure of an operator to be self-adjoint is related to boundary conditions

on a non-compact manifold M . We will make this more precise soon, but some technical results are required.

Elements of the proof of the following result were adapted from [10, Proposition 3.2].

Lemma 2.11. Let K ⊂ M be compact. For sufficiently small t > 0, there exist operators Ft ∶ L2(K,E) →

L2(M,E) such that

1. ∥Ft∥ < 1,

2. for each u ∈ L2(K,E), Ftu→ u in L2(M,E) as t→ 0,

3. for each u ∈ L2(K,E), Ftu is smooth, with compact support, and

4. the commutator [D,Ft] extends to a bounded operator from L2(K,E) to L2(M,E), whose norm is

bounded independent of t.

We call such a family Friedrichs mollifiers on M .

6

Page 8: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Proof. Fix ε > 0. Let ϕ ∶ Rn → R be a smooth positive function with compact support and total mass 1

with supp(ϕ) ⊆ B1(0). There are many choices of such a function, one of which is ϕ(x) = cne− 1

1−∥x∥2 on

B1(0) extended to 0 on Rn, where cn is a normalising constant. Define Ft ∶ L2(Rn) → L2(Rn) by Ftu(x) =

t−n ∫Rnu(y)ϕ(x − y

t)dy. Note that Ftu(x) = (u ⋆ ϕt)(x), where ϕt(x) = t−nϕ(x

t).

To estimate the norm, we apply Young’s inequality to see that

∥Ftu∥2 = ∥u ⋆ ϕt∥2

≤ ∥u∥2 ∥ϕt∥1

= ∥u∥2 .

Hence taking the supremum over all u with ∥u∥2 = 1, we have that ∥Ft∥ ≤ 1.

Next, we aim to show L2 convergence. To this end, we first show that Ftw → w uniformly for w ∈ Cc(Rn).

Note that supp(ϕt) ⊆ Bt(0) and so we may find s > 0 such that sup{∣w(x − y) −w(x)∣ ∶ ∣y∣ ≤ t} < ε for t < s. So

fix t < s and calculate that

∣Ftw(x) −w(x)∣ = ∣∫Rnw(x − y)ϕt(y)dy − ∫

Rnw(x)ϕt(y)dy∣

≤ ∫Rn

∣w(x − y) −w(x)∣ϕt(y)dy

≤ sup{∣w(x − y) −w(x)∣ ∶ ∣y∣ ≤ t}

< ε.

Find v ∈ Cc(Rn) such that ∥u − v∥ < ε3

. Applying property 1 we have that

∥Ftu − Ftv)∥2 = ∥Ft(u − v∥2

≤ ∥u − v∥2

< ε3.

Choose t > 0 such that ∥Ftv − v∥2 <ε

3(possible since uniform convergence implies L2 convergence) and calculate

that

∥Ftu − u∥2 ≤ ∥Ftu − Ftv∥2 + ∥Ftv − v∥2 + ∥u − v∥2

≤ 2ε

3+ ∥Ftv − v∥

< ε.

Hence Ftu→ u in L2.

By definition, we have that supp(Ftu) ⊆ {x + y ∶ x ∈ supp(u) and y ∈ Bt(0)} which is bounded. Hence

7

Page 9: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

supp(Ftu) is compact. To check smoothness, we calculate that

limh→0

Ftu(x + hei) − Ftu(x)h

= limh→0∫Rn

ϕt(x + hei − y) − ϕth

(x − y)u(y)dy

= ∫Rn

∂ϕt(x − y)∂xi

u(y)dy

= u ⋆ ∂ϕ

∂xi.

So we see inductively that Ftu is smooth, since ϕt is smooth. Note also that in a similar manner to the above

calculation, we can show that∂

∂xj(u ⋆ ϕt) = ∂u

∂xj⋆ ϕt = u ⋆

∂ϕt∂xj

. So we calculate the commutator for any

u ∈ C∞c (Rn) as

[D,Ft]u(x) =D(Ftu) − Ft(Du)

=∑j

Aj∂

∂xj(u ⋆ ϕt) +B(u ⋆ ϕt) −∑

j

(Aj∂u

∂xj) ⋆ ϕt − (Bu ⋆ ϕt)

=∑j

Aj (u ⋆∂ϕt∂xj

) +B(u ⋆ ϕt) −∑j

(Aj∂u

∂xj) ⋆ ϕt − (Bu ⋆ ϕt),

where we have moved the derivative from u to ϕ as discussed before the calculation. We now write the

convolution in integral form and apply integration by parts, giving

[D,Ft]u(x) =∑j

(Aj(x)∫Rnu(y)∂ϕt

∂xj(x − y)dy − ∫

RnAj(y)

∂u

∂yjϕt(x − y)dy)

+B(x)∫Rnu(y)ϕt(x − y)dy − ∫

RnB(y)u(y)ϕt(x − y)dy

=∑j∫Rn

[(Aj(x) −Aj(y))u(y)∂ϕt∂xj

(x − y) −Aj(y)∂

∂yj(u(y)ϕt(x − y))]dy

+ ∫Rn

(B(x) −B(y))u(y)ϕt(x − y)dy.

We now rearrange terms and use the fact that∂ϕt∂xj

(x − y) = −∂ϕt∂yj

(x − y) to obtain

[D,Ft]u(x) =∑j∫Rn

(Aj(x) −Aj(y))u(y)∂ϕt∂xj

(x − y)dy + ∫Rn

(B(x) −B(y))u(y)ϕt(x − y)dy

=∑j

1

t∫Rn

(Aj(x) −Aj(y))u(y)∂ϕ

∂xj(x − y

t)dy + ∫

Rn(B(x) −B(y))u(y)ϕt(x − y)dy.

Using local coordinates and a partition of unity, it is possible to graft this onto an arbitrary manifold to

construct a family of Friedrichs mollifiers there. The proof and technique for this are quite involved and will

not be presented here.

We will need another technical lemma before being able to prove the result about ‘boundary conditions’,

however the proof is omitted. The proof can be found in [5, Lemma 1.8.1].

Lemma 2.12. Let T be a closable operator on a Hilbert space H. Then u ∈H belongs to the minimal domain

if and only there is a sequence (uj) in the domain of T such that uj → u while ∥Tuj∥ remains bounded.

8

Page 10: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Lemma 2.13. Let D be a symmetric differential operator on a bundle E over a manifold M and suppose that

u is a compactly supported element of L2(M,E). Then u belongs to the minimal domain of D if and only if it

belongs to the maximal domain.

Proof. We use the Friedrichs’ mollifiers developed in Lemma 2.11. Suppose u ∈ Dom(D∗) is supported in K.

We know that Ftu→ u as t→ 0 and Ftu is compactly supported and smooth. We also have that

DFtu = FtDu + [D,Ft]u,

which is bounded in norm by Lemma 2.11. Hence by Lemma 2.12, we have that u belongs to the maximal

domain if and only if belongs to the minimal domain.

Corollary 2.14. Every compactly supported symmetric differential operator on an open manifold is essentially

self-adjoint.

Proof. Applying Lemma 2.13, we see that Dom(D) = Dom(D∗) = Dom(D∗). Hence D is self-adjoint and D is

essentially self-adjoint.

3 Algebraic Prerequisites

We wish to work with a specific class of differential operators, called Dirac operators. However before defining

Dirac operators, we must first develop some algebraic prerequisites from the theory of Clifford algebras and

vector bundles. The material in this section is primarily found in [7, Chapter I].

3.1 Clifford Algebras

Definition 3.1. Let V be a finite dimensional vector space and q ∶ V ×V → C a non-degenerate inner product.

Define an ideal Jq in the tensor algebra T (V ) =⊕n≥0

V ⊗n generated by elements of the form v ⊗ v + q(v, v)1 for

v ∈ V . We define the Clifford algebra Cliff(V, q) to be the quotient

Cliff(V, q) = T (V )/Jq.

Note that Cliff(V, q) is generated by 1 and v,w ∈ V satisfying

v ⋅w +w ⋅ v = −2q(v,w).

Hence if v and w are orthogonal they anticommute in Cliff(V, q). Clifford algebras also satisfy a universal

property as follows.

Lemma 3.2. If A is a unital associative algebra and c ∶ V → A is a linear map satisfying

c(v)c(v) = −q(v, v)1A

for all v ∈ V , there exists a unique algebra homomorphism c ∶ Cliff(V, q)→ A extending c.

9

Page 11: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Proof. Firstly, there exists a unique algebra homomorphism ϕ ∶ T (V )→ A such that ϕ○ι = c, where ι ∶ V → T (V )

is the inclusion map, defined by ϕ(v1 ⊗⋯⊗ vn) = c(v1)⋯c(vn). Next, check that for v ∈ V we have

ϕ(v ⊗ v + q(v, v)) = ϕ(v ⊗ v) + ϕ(q(v, v)

= c(v) ⋅ c(v) + q(v, v)1A

= 0.

Hence ϕ vanishes on Jq and so descends uniquely to c on Cliff(V, q).

Example 3.3. Note that if A ∈ O(V, q) = {A ∶ V → V ∶ q(Av,Av) = q(v, v) for all v ∈ V }, then we may define a

linear map c ∶ V → Cliff(V, q) by c(v)w = Av ⋅w. Since this satisfies

c(v)c(v) = Av ⋅Av

= −q(Av,Av)

= −q(v, v).

Hence by Lemma 3.1, c extends uniquely to an automorphism c ∶ Cliff(V, q) → Cliff(V, q). If we consider the

map α ∶ V → V defined by α(v) = −v, we have an involution on Cliff(V, q) and thus we may decompose into the

eigenspaces for ±1 as

Cliff(V, q) = Cliff0(V, q)⊕Cliff1(V, q).

Since α(v1 ⋅ ⋅ ⋅ ⋅ ⋅ vn) = (−1)nv1⋯vn we see that Cliff(V, q) decomposes into sums of even and odd products of

vectors. Since α is an algebra homomorphism, we see that

Cliffi(V, q) ⋅Cliffj(V, q) ⊆ Cliffi+j(V, q),

where the addition of indices is mod 2. Algebras satisfying this property are said to be Z2-graded.

3.2 Clifford Modules and Bundles

Definition 3.4. A C-representation of the Clifford algebra Cliff(V, q) is an R-algebra homomorphism ϕ ∶

Cliff(V, q) → EndC(W ) into the algebra of complex linear transformations of a finite dimensional complex

vector space W . A vector space W together with a Cliff(V, q) representation is called a Cliff(V, q) module.

For v ∈ Cliff(V, q) and w ∈W , we write

v ⋅w = ϕ(v)(w)

and refer to this as Clifford multiplication of w by v. We may define R-representations and H-representations

in a similar manner.

10

Page 12: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Definition 3.5. A C-representation is called reducible if there exists a decomposition W =W1 ⊕W2 such that

ϕ(v)(Wj) ⊂ Wj for j = 1,2 and all v ∈ Cliff(V, q). In this case we may write ϕ = ϕ1 ⊕ ϕ2. A representation is

irreducible if there are no proper invariant subspaces. Two representations are equivalent if there exists a C

linear isomorphism F ∶W1 →W2 such that Fϕ1(v)F −1 = ϕ2(v) for all v ∈ Cliff(V, q).

It is a non trivial fact that every complex Cliff(V, q) module decomposes uniquely as a sum of irreducibles.

Example 3.6. Define ϕ ∶ Cliff(V, q) → Cliff(V, q) by ϕ(v)w = v ⋅ w. Then we see that Cliff(V, q) is trivially a

Cliff(V, q) module.

Example 3.7. For a more involved example, consider the exterior algebra Λ∗V . Let (v1, . . . , vn) be an ortho-

normal basis for V . Define exterior multiplication by vj as ej and interior multiplication by vj as ij . Recall

that ejik + ikej = δjk. Define ϕj = ej − ij and calculate that

ϕj ○ ϕk + ϕk ○ ϕj = (ej − ij)(ek − ik) + (ek − ik)(ej − ij)

= ejek − ejik − ijek + ijik + ekej − ekij − ikej + ikij

= −2δjk.

So for a fixed vj , define a linear map ϕ ∶ V → End(Λ∗V ) by ϕ(v) = ϕj(v). We can uniquely lift this to

ϕ ∶ Cliff(V )→ End(ΛV ) by the universal property (Lemma 3.1). Hence Λ∗V is a Cliff(V, q) module.

We now wish to extend the properties of Clifford algebras to vector bundles, in order to apply the results to

manifolds and their tangent and cotangent bundles.

Recall that by Example 3.3, for any A ∈ SOn we have a unique extension to Cliff(Rn). Hence we have a

representation ρn ∶ SOn → Aut(Cliff(Rn)) defined by lifting the transformation to the Clifford algebra. This

representation will allow us to construct bundles of Clifford algebras.

Definition 3.8. The Clifford bundle of the oriented Riemannian vector bundle E over the manifold M is the

bundle

Cliff(E) = PSO(S) ×ρn Cliff(Rn).

Note that we could also have defined Cliff(E) as the quotient bundle

Cliff(E) = T (E)/I(E),

where T (E) = ⊕n≥0

E⊗n and I(E) is the bundle of ideals whose fibre at x ∈ M is the two-sided ideal I(Ex) in

T (Ex) generated by elements v ⊗ v + ∥v∥2 for v ∈ Ex. From this definition we see that Cliff(E) is a bundle of

Clifford algebras over M . By defining fibrewise multiplication for sections u, v ∶M → Cliff(E) as

(u ⋅ v)x = u(x)v(x)

11

Page 13: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

we may give an algebra structure to the space of sections, Γ(E). Note that many notions for Clifford algebras

carry through to the theory of Clifford bundles. For example, if we consider the bundle automorphism

α ∶ Cliff(E)→ Cliff(E)

defined by extending the map α in Example 3.1, then the eigenbundles for ±1 gives a corresponding decompo-

sition

Cliff(E) = Cliff0(E)⊕Cliff1(E).

3.3 Spinor Bundles

Definition 3.9. The spin groups are defined, for n ∈ N, as

Spin(n) = {v1⋯vk ∶ vi ∈ Rn, ∥vi∥ = 1 for 1 ≤ i ≤ k and k is even}.

Note that for a unit vector v ∈ Rn we have v ⋅ v = −1 and v ⋅ (−v) = 1 we see that {1,−1} ⊂ Spin(n). We also see

that by definition Spin(n) ⊆ Cliff0(Rn).

For n ≥ 3 there exists a universal covering homomorphism ξ0 ∶ Spin(n) → SO(n). The kernel of this

homomorphism is {−1,1}.

Definition 3.10. Suppose n ≥ 3. Then a spin structure on E is a principal Spin(n) bundle PSpin(E) together

with a 2-sheeted covering

ξ ∶ PSpin(E)→ PSO(E)

such that ξ(pg) = ξ(p)ξ0(g) for all p ∈ PSpin(E) and all g ∈ Spin(n).

When n = 2, the definition remains the same except that ξ0 ∶ SO2 → SO2 is now the connected double cover.

When n = 1, PSO(E) ≅M and a spin structure is a double cover of M .

Definition 3.11. Let E →M be an oriented Riemannian vector bundle with a spin structure ξ ∶ PSpin(E) →

PSO(E). A real spinor bundle of E is a bundle of the form

S(E) = PSpin(E) ×µ L,

where L is a left module for Cliff(Rn) and µ ∶ Spinn → SO(L) is the representation given by left multiplication

by elements of Spinn ⊂ Cliff0(Rn).

Likewise, a complex spinor bundle of E is a bundle of the form

SC(E) = PSpin(E) ×µ LC,

where LC is a complex left module for Cliff(Rn)⊗C. If L (or LC) is Z2-graded, we say the corresponding spinor

bundle is Z2-graded.

Example 3.12. Consider Cliff(Rn) as a module acting on itself by left multiplication µ. This gives a ’principal

Cliff(Rn) bundle‘ with a free right action α ∶ Cliff(Rn) × S(Cliff(Rn))→ S(Cliff(Rn)).

12

Page 14: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

3.4 Connections

Definition 3.13. Let M be a Riemannian manifold and let E be a smooth (R or C) vector bundle over M of

dimension n. A connection on E is a linear map ∇ ∶ C∞(M,E)→ C∞(M,T ∗M ⊗E) such that

∇(fu) = df ⊗ u + f∇u

for all f ∈ C∞(M) and u ∈ C∞(M,E). Given a smooth vector field X ∶ M → TM we thus have a map

∇X ∶ C∞(M,E) → C∞(M,E), called the covariant derivative with respect to X. Note that this covariant

derivative satisfies

∇X(fu) =X(f)u + f∇Xu.

Definition 3.14. Let (v1, . . . , vn) be a local orthonormal basis for Ex and (e1, . . . , em) a basis for TxM . Then

we define the connection coefficients by

∇eivj = Γkijvk.

Definition 3.15. A connection ∇ on a Riemannian bundle E is called Riemannian if it satisfies the inner

product rule

X⟨u,u′⟩ = ⟨∇Xu,u′⟩ + ⟨u,∇Xu′⟩

for all X ∈ C∞(M,TM) and u,u′ ∈ C∞(M,S).

4 Dirac Operators

We give a motivation for the development of Dirac operators as presented in [7, Chapter 2] The study of Dirac

operators begins historically with physicist Paul Dirac, who was searching for a first order differential operator

that was Lorentz invariant and squared to the Klein-Gordon operator,

1

c2∂2

∂t2−

3

∑j=1

∂2

∂x2j+ m

2c2

h2.

Thus he was essentially looking for an operator D satisfying D2 = ◻ (the d’Alembertian) compatible with special

relativity, and guessed the form

D =n

∑k=1

ψk∂

∂xk.

Dirac realised that the ψk must be matrices and satisfy what we now recognise as the generating relations for

a representation of Cliff(Rn). Generalisations of these operators will be the differential operators of interest to

us.

In modern physics we are often working in 3D-spacetime, called Minkowski space, rather than Euclidean

space. Minkowski space is defined as the inner product space (Rn+1, q) where for x = (x0, x1, . . . , xn) and

13

Page 15: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

y = (y0, y1, . . . , yn) in Rn we have

q(x, y) = −x0y0 +n

∑k=1

xkyk.

This is the framework used for the study of special relativity. Just as Euclidean space can be generalised

to Riemannian manifolds, Minkowski space can be generalised to an object called a Lorentzian manifold. A

Lorentzian manifold is defined in the same manner as a Riemannian manifold however the metric is no longer

required to be positive definite, but is required to have signature (1, n) where (n + 1) is the dimension of the

manifold. For an in depth discussion on pseudo-Riemannian and Lorentzian geometry, see [8] and [1, Appendix

2].

4.1 Definition

Definition 4.1. Let M be a Riemannian manifold of dimension n and E a smooth complex vector bundle

over M . Write left Clifford multiplication as c ∶ C∞(M,TM ⊗ E) → C∞(M,E) and let ∇ ∶ C∞(M,E) →

C∞(M,T ∗M ⊗ E) be a connection. Let φ ∶ C∞(M,T ∗M ⊗ E) → C∞(M,TM ⊗ E) denote the isomorphism

between vector and covector fields. We define the Dirac operator associated to the bundle E and connection ∇,

D ∶ C∞(M,E)→ C∞(M,E), as D = c ○ φ ○ ∇. Note that this also depends on the representation chosen.

Thus for an orthonormal frame (ek) of (TM)x, with dual frame (ek) of (T ∗M)x we have that

Du(x) = c ○ φ(n

∑k=1

ek ⊗∇eku∣x)

= c(n

∑k=1

ek ⊗∇eku∣x)

=n

∑k=1

ek ⋅ ∇eku∣x.

Example 4.2. Choosing normal coordinates at x ∈ M so that∂

∂xj(x) = ej(x), we may compute the Dirac

14

Page 16: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Laplacian D2 as

D2u(x) =n

∑k=1

ek(x) ⋅ ∇ek(x) (Du)

=n

∑k=1

ek(x) ⋅ ∇ek(x)⎛⎝

n

∑j=1

ej(x) ⋅ ∇ej(x)u⎞⎠

=n

∑k=1

n

∑j=1

[ek(x) ⋅ ej(x) ⋅ ∇ek(x)∇ek(x)u + ek(x)∇ek(x)ej(x) ⋅ ∇ej(x)u]

=n

∑k=1

n

∑j=1

[−gjk ⋅ ∇ek(x)∇ej(x)u + ek(x)∇ek(x)ej(x) ⋅ ∇ej(x)u]

=n

∑k=1

n

∑j=1

[−gjk ⋅ ∇ ∂

∂xk∇ ∂

∂xjulfl + ek(x)∇ek(x)ej(x) ⋅ ∇ ∂

∂xju]

=n

∑k=1

n

∑j=1

[−gjk ⋅ (∇ ∂

∂xj( ∂ul

∂ej(x)fl) + ul∇ ∂

∂xjfl) + ek(x)Γlkjfl ⋅ ∇ ∂

∂xj(ulfl)]

=n

∑k=1

n

∑j=1

[−gjk ⋅ ( ∂2ul

∂xkxjel +

∂ul

∂xj∇ ∂

∂xjvl + ulΓmjlvm) + ek(x)Γmkjvm ⋅ ( ∂u

l

∂xjvl + ul∇ ∂

∂xj)vl)]

=n

∑k=1

n

∑j=1

[−gjk ⋅ ( ∂2ul

∂xkxjvl +

∂ul

∂xjΓmklvm + ek(x)ulΓmjlvm) + Γmkjvm ⋅ ( ∂u

l

∂xjvl + ulΓpjlvp)]. (4.1)

Note that everything in this computation is evaluated at x ∈M . From this computation we see that

σD2(x, ξ) = −n

∑j=1

n

∑k=1

gjkξjξk

= − ∥ξ∥2 .

Thus D2 is a Laplace type operator, as expected.

4.2 Examples

We will now examine some Euclidean examples of Dirac operators. Consider the simple manifold M = Rn with

the standard inner product and let Cliff(Rn) be it’s Clifford algebra. Let W be a complex vector space of

dimension m and ϕ ∶ Cliff(Rn) → EndC(W ) be a representation of Cliff(Rn). Let E =M ×W have the trivial

Levi-Civita connection. We consider the Dirac operator D ∶ C∞(M,E)→ C∞(M,E). Then from Definition 4.1

we have for u ∈ C∞(M,E) and {e1, . . . , en} the standard basis vectors for Rn that

Du =n

∑k=1

ek ⋅ ∇eku

=n

∑k=1

ϕ(ek)∂u

∂xk

=n

∑k=1

ϕk∂u

∂xk

where ϕ(ek) ∶= ϕk ∶W →W are linear maps satisfying

ϕk ○ ϕj + ϕj ○ ϕk = −2δkjIdW . (4.2)

15

Page 17: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

So identifying EndC(W,W ) ≅ Mm(C) and applying Equation 4.2 with j = k shows that ϕk ∈ GLm(C) for all

1 ≤ k ≤ n. Note also that applying 4.1 and the fact that gjk = δjk and Γkij = 0 for all i, j, k gives

D2u = −n

∑k=1

∂2u

∂xkIdW .

Example 4.3. Let n = 1, so Cliff(R) = C and let W = C. Then we have a single linear ϕ ∶ C → C satisfying

ϕ ○ ϕ = −1. So we have two choices, ϕ(x) = ±ix. The usual convention is ϕ(x) = −ix for the Dirac operator to

be self adjoint. Thus the Dirac operator on R is

D = 1

i

d

dx.

Example 4.4. Let n = 2 and W = C ⊕C, so that Cliff(R2) = H ≅ C ⊕C, as R-linear spaces (not as algebras).

Note that the decomposition of H into C⊕C corresponds to the Z2 grading Cliff(R2) = Cliff0(R2)⊕Cliff1(R2).

To see this, define α1 ∶ Cliff0(R2) → C and α2 ∶ Cliff1(R2) by α0(x + ye1e2) = x + iy = α2(xe1 + ye2). Define

ϕ0 = ϕ(1) = Id. To find ϕ1 and ϕ2, we note that we need two 2×2 matrices who square to −Id and anticommute.

Hence we recall two of the three Pauli matrices as

ϕ1 =⎛⎜⎝

0 −1

1 0

⎞⎟⎠

and

ϕ2 =⎛⎜⎝

0 i

i 0

⎞⎟⎠.

Note that the third Pauli matrix is generated by the relation

ϕ(e1e2) = ϕ(e1)ϕ(e2)

=⎛⎜⎝

−i 0

0 i

⎞⎟⎠.

Then we have for u ∈ C∞(R2, S) that

Du = ϕ1∂u

∂x1+ ϕ2

∂u

∂x2

=⎛⎜⎝

0 − ∂u∂x1 + i ∂u∂x2

∂u∂x1 + i ∂u∂x2 0

⎞⎟⎠

=⎛⎜⎝

0 −∂u∂z

∂u∂z

0

⎞⎟⎠.

Thus we see that on R2, D is the Cauchy-Riemann operator.

A more interesting example to consider is the Dirac operator on a cylinder.

Example 4.5. Let M = R × S1 be the infinite cylinder with the product metric g = dt2 + dθ2. Let C =

[t1, t2] × S1 ⊂M be the finite cylinder with metric g = g∣C and S = C ×C2. The Levi-Civita connection on C is

16

Page 18: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

flat. Note that Cliff(T ∗M,g) = H and we aim to find a representation for this algebra. So we aim to find an R

algebra homomorphism ϕ ∶ H→ EndC(C2). Define ϕ ∶ H→ EndC(C2) by

ϕ(α + iβ + jγ + kδ) = α⎛⎜⎝

1 0

0 1

⎞⎟⎠+ β

⎛⎜⎝

0 −1

1 0

⎞⎟⎠+ γ

⎛⎜⎝

0 i

i 0

⎞⎟⎠+ δ

⎛⎜⎝

i 0

0 −i

⎞⎟⎠.

Note that a basis for Cliff(T ∗M) is {1,dt,dθ,dtdθ} and so we see that we are identifying ϕ(dt) =⎛⎜⎝

0 −1

1 0

⎞⎟⎠

and

ϕ(dθ) =⎛⎜⎝

0 i

i 0

⎞⎟⎠

.

So applying Definition 4.1 we have for a local basis {v1, v2} of C that

Du =2

∑j=1

2

∑l=1

∂xj⋅ ∇ ∂

∂xj(ulvl)

=2

∑j=1

2

∑l=1

∇ ∂

∂xj( ∂u

l

∂xjvl)

=2

∑j=1

2

∑l=1

dxj ( ∂ul

∂xjvl)

=2

∑l=1

(dt(∂ul

∂tvl) + dθ (∂u

l

∂θvl))

=⎛⎜⎝

0 −∂tu2

∂tu1 0

⎞⎟⎠+⎛⎜⎝

0 i∂θu2

i∂θu1 0

⎞⎟⎠

=⎛⎜⎝

0 −∂t + i∂θ∂t + i∂θ 0

⎞⎟⎠

⎛⎜⎝

u1

u2

⎞⎟⎠.

Note that this has the same form as the Dirac operator on R2 in Example 4.2.

5 A Dirac Operator on the Lorentzian Cylinder R × S1

In physical problems, we are often working with a Lorentzian manifold. We will now consider the Dirac operator

on a cylinder with a Lorentzian metric. We will impose APS boundary conditions as found in [2] and solve

the eigenvalue problem for D. The aim of this section is to apply the results of [2] to a simple situation and

compute some traces with the goal of combining this work with that presented by Connes’ in [3], generalising

to the setting of Lorentzian rather than Riemannian manifolds.

5.1 Constructing the Dirac Operator

Let M = R×S1 be the infinite cylinder with the Lorentzian metric g = −dt2+dθ2. Let C = [t1, t2]×S1 ⊂M be the

finite cylinder with metric g = g∣C and S = C ×C2 be a spinor bundle over C. The Levi-Civita connection on C

is flat. Note that since the metric is Lorentzian, D will no longer be an elliptic operator, rather a generalisation

of a hyperbolic operator. We use the following definition from [1].

17

Page 19: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Definition 5.1. Let M be a Lorentzian manifold and let E →M be a real or complex vector bundle. A second

order linear differential operator D ∶ C∞(M,E) → C∞(M,E) is called normally hyperbolic if its principal

symbol is given by the metric,

σD(x, ξ) = − ∥ξ∥2 ⋅ IdEx ,

for all (x, ξ) ∈ T ∗M .

Example 5.2. The computation in Example 4.2 works for a metric of any signature, hence for a Dirac operator

D on a Lorentzian manifold we have that D2 is a normally hyperbolic operator rather than an elliptic operator.

Note also that Cliff(T ∗M,g) = M2(R) and we aim to find a representation for this algebra. So we aim to

find an R-algebra homomorphism ϕ ∶M2(R)→ EndC(C2).

Construct a basis of M2(R) as

⎧⎪⎪⎪⎨⎪⎪⎪⎩Id =

⎛⎜⎝

1 0

0 1

⎞⎟⎠, e1 =

⎛⎜⎝

0 1

1 0

⎞⎟⎠, e2 =

⎛⎜⎝

0 1

−1 0

⎞⎟⎠, e1e2 = e3 =

⎛⎜⎝

−1 0

0 1

⎞⎟⎠

⎫⎪⎪⎪⎬⎪⎪⎪⎭.

Note that this basis satisfies

Id2 = e21 = −e22 = e23 = 1.

So we define ϕ ∶ M2(R) → EndC(C2) by inclusion, which is clearly an R-algebra homomorphism. So we

identify dt = e1 and dθ = e2 as in the previous example.

So applying Definition 4.1 we have for a local basis {v1, v2} of C that

Dψ =2

∑j=1

2

∑l=1

∂xj⋅ ∇ ∂

∂xj(ψlvl)

=2

∑j=1

2

∑l=1

∇ ∂

∂xj(∂ψ

l

∂xjvl)

=2

∑j=1

2

∑l=1

dxj (∂ψl

∂xjvl)

=2

∑j=1

dxj∂ψ

∂xj

=⎛⎜⎝

0 ∂ψ∂t

∂ψ∂t

0

⎞⎟⎠+⎛⎜⎝

0 −∂ψ∂θ

∂ψ∂θ

0

⎞⎟⎠

=⎛⎜⎝

0 ∂ψ∂t

− ∂ψ∂θ

∂ψ∂t

+ ∂ψ∂θ

0

⎞⎟⎠.

5.2 APS Boundary Conditions

As is usual with physical problems, we would like to impose boundary conditions on our operator D. The

boundary conditions we will work with are Atiyah-Patodi-Singer (APS) boundary conditions, as described in

18

Page 20: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

[2]. We define the APS spaces as

C∞APS(C,S+) = {ψ ∈ C∞(C,S+) ∶ P≥0(t1)ψ = 0 = P≤0(t2)ψ},

C∞APS(C,S−) = {ψ ∈ C∞(C,S−) ∶ P≤0(t1)ψ = 0 = P≥0(t2)ψ}.

These definitions were adapted from those given in [2, Section 1.3]. Here we have P≥0(t) = χ[0,∞)(∂θ ∣t) and

P≤0(t) = χ(−∞,0](∂θ ∣t), spectral projections onto the non-negative and non-positive eigenspaces for the operator

∂θ respectively. Physically, C∞APS(C,S+) represents particles whose energy is non-positive at time t1 and is

non-negative at time t2. Likewise for C∞APS(C,S−). To see what this means explicitly, first note that we can

decompose any function ψ ∶ C → C in a Fourier basis as

ψ(t, θ) = ∑n∈Z

ψn(t, θ).

A particle has entirely non-positive energy at t1 if ψn(t1, θ) = 0 for all n ≥ 0 and non-negative energy at time t2

if ψn(t2, θ) = 0 for all n ≤ 0. Hence we find that

D =⎛⎜⎝

0 D−

D+ 0

⎞⎟⎠

=⎛⎜⎝

0 ∂t − iA

∂t + iA 0

⎞⎟⎠,

where A = 1

i∂θ, D+ = ∂t + iA and D− = ∂t − iA. Note also that Dom(D+) = C∞

APS(C,S+) and Dom(D−) =

C∞APS(C,S−).

We consider the Dirac operator acting on spinors ψ which take the form ψ =⎛⎜⎝

ψ+

ψ−

⎞⎟⎠

, where ψ+ represents a

particle and ψ− represents an antiparticle.

5.3 Determining the Eigenvalues

As is standard when working with Dirac operators, we first find the eigenvalues of the associated Laplacian.

Compute that

D2 =⎛⎜⎝

D−D+ 0

0 D−D+

⎞⎟⎠

=⎛⎜⎝

∂2t − ∂2θ 0

0 ∂2t − ∂2θ

⎞⎟⎠.

Note that although D+D− and D−D+ have the same expression, they are in fact different operators, since

Dom(D+D−) = {ψ ∈ Dom(D−) ∶D−ψ ∈ Dom(D+)} and Dom(D−D+) = {ψ ∈ Dom(D+) ∶D+ψ ∈ Dom(D−)}.

As mentioned, a solution ψ will be particle-antiparticle pair of the form ψ =⎛⎜⎝

ψ+

ψ−

⎞⎟⎠

. As usual with a linear

PDE, we apply the method of separation of variables. First, suppose ψ+(t, θ) = T (t)Θ(θ). Applying this to the

19

Page 21: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

eigenvalue problem we see that

T ′′(t)T (t)

− Θ′′(θ)Θ(θ)

= λ2.

So we have that Θ′′(θ) = −kΘ(θ) and T ′′(t) = (−k+λ2)T (t) for some constant k ∈ C. Since Θ(θ) ∈ S1 we require

that Θ(0) = Θ(2mπ) for all m ∈ Z and so this has solution if and only if√k ∈ Z. Hence for each n ∈ Z we have

the solution Θn(θ) = einθ. Next, we see that for each n ∈ Z solutions Tn(t) will be of the form

Tn(t) = αnei√n2−λ2(t−t1) + βne−i

√n2−λ2(t−t1)

for some αn, βn ∈ C when λ2 ≠ n2 and

Tn(t) = αn(t − t1) + βn

when λ2 = n2. Suppose for now that λ2 ∈ R.

We now apply our APS boundary conditions to determine possible values of λ. The APS boundary conditions

tell us that Tn(t1) = 0 for n ≥ 0 and Tn(t2) = 0 for n ≤ 0. So for n ≥ 0 and n2 ≠ λ2 we have αn = −βn and for

n ≥ 0 and n2 = λ2 we have βn = 0. For n ≤ 0 we have Tn(t2) = 0. This forces αn = −e−2i√n2−λ2(t2−t1)βn when

n2 ≠ λ2 and αn = 0 when n2 = λ2. Hence we have eigenvectors

ψ+n,λ(t, θ) =

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎩

(ei√n2−λ2(t−t1) − e−i

√n2−λ2(t−t1)) einθ, if n ≥ 0 and λ2 ≠ n2,

(ei√n2−λ2(t−t1) − e−i

√n2−λ2(t−2t2+t1)) einθ, if n ≤ 0 and λ2 ≠ n2,

0, if λ2 = n2.

We find similarly for antiparticles that

ψ−n,λ(t, θ) =

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎩

(ei√n2−λ2(t−t1) − e−i

√n2−λ2(t−t1)) einθ, if n ≤ 0 and λ2 ≠ n2,

(ei√n2−λ2(t−t1) − e−i

√n2−λ2(t−2t2+t1)) einθ, if n ≥ 0 and λ2 ≠ n2,

0, if λ2 = n2.

= ψ+−n,λ(t, θ).

Now we attempt to solve the eigenvalue problem

(∂t + ∂θ)ψ+n,λ = λψ−n,λ.

With n ≥ 0 and n2 ≠ λ2 this leads to

0 = [(i√n2 − λ2 + in − λ) ei

√n2−λ2(t−t1) + (i

√n2 − λ2 − in + λe2i

√n2−λ2(t2−t1))] einθ.

Comparing coefficients and using linear independence we see that

i√n2 − λ2 + in = λ

20

Page 22: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

and

i√n2 − λ2 − in = −λe2i

√n2−λ2(t2−t1).

Combining these gives

n +√n2 − λ2

n −√n2 − λ2

= −e2i√n2−λ2(t2−t1). (5.1)

If n2 > λ2 then the left hand side is real, forcing

2√n2 − λ2(t2 − t1) =mπ

for some m ∈ Z. Rearranging we see that the eigenvalues satisfy

λ2 = n2 − m2π2

4(t2 − t1)2.

If n2 < λ2, let√n2 − λ2 = iρn where ρn ∈ R and so

n + iρnn − iρn

= e2ρn(t2−t1).

The left hand side is real and so we must have ρn = 0, a contradiction. A similar analysis for n ≤ 0 gives the

same result. Hence we have eigenvalues

λn,m =

¿ÁÁÀn2 − m2π2

4(t2 − t1)2,

where n,m ∈ Z and n2 > m2π2

4(t2 − t1)2. From this expression we see that finding an analogue of Weyl’s theorem in

this case will be difficult, as the eigenvalues are indexed by two variables.

Next, consider the case λ2 ∉ R. Then we must have λ = α + iβ for some 0 ≠ α,β ∈ R. Note that all of the

analysis up to and including Equation 5.1 still applies.

5.4 Action Principles

Many physical problems require us to find and solve equations of motion. To do this, a common technique is to

develop an action principle, which usually involves finding minima of functions called actions. For an excellent

overview of material regarding action principles in physics, refer to [11].

Example 5.3 (Fermat’s Principle). In optics, Fermat’s principle states that light will take the path of shortest

time. Mathematically, this says that light will minimise the integral

∫t2

t1

ds

v.

Example 5.4 (Hamilton’s Principle). In classical mechanics, Hamilton’s principle states that the actual path

taken by a particle will minimise the action of the Lagrangian, that is it will minimise the integral

S(x, x) = ∫t2

t1L(t, x(t), x(t))dt,

where L = T −U is the Lagrangian, T is the kinetic energy and U is the potential energy.

21

Page 23: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Physicist Richard Feynman attempted to extend this to quantum mechanics with his path integral formali-

sation, however this is often poorly defined and difficult to work with, so another approach is needed.

5.5 Connes’ Spectral Action Principle

In the 1990’s, Connes’ attempted to develop an action principle for compact Riemannian manifolds defined in

terms of a trace of a differential operator, found in [3].

Let H be a Hilbert space, A an algebra of bounded operators on H and D a self-adjoint operator on H.

Consider an operator on H of the form DA =D+A+JAJ∗, where J is an anti-linear isometry. Connes’ spectral

spectral action principle is define in terms of the trae

Trace(χ(D2A

Λ2)) ,

where χ ∶ σ(DA) → R is measurable and satisfies (χ(λ) = 1 for λ ≤ 1, and Λ ≥ 0 is a spectral cutoff. For large

values of Λ this can be computed using a heat kernel expansion, with the coefficients containing information

regarding the geometry.

However spacetime is neither compact nor Riemannian, so more work is required for this to apply to a non-

idealistic physics problem. Compactness can be overcome by multiplying by a compactly supported function,

however it is not so simple to move to the Lorentzian setting. To do this, we would like to be able to define a

trace on such Dirac operators D.

5.6 Existence and Analysis of Trace Functionals

Before being able to determine any useful geometric or physical information from these traces, we must first do

some analysis to determine that the integrals involved converge. In this section we present some candidates for

a trace and some calculations involving them.

A simple trace we would like to compute is

Trace (Mf (1 +D2APS,ε,R)

− s2 ) ,

where we have imposed an energy cutoff R and a lower limit ε > 0 for our operator. To compute this, we will

require some technical results.

Lemma 5.5. Let f ∶ Rn → C be rotationally and reflectionally symmetric. Then for any v ∈ Rn with ∥v∥ = 1

we have that

∫Rnf(x)dx = Vol (Sn−1)∫

0f(rv)rn−1 dr.

Proof. First, recall the form of the Beta function β(x, y) = 2∫π2

0sin2x−1 (θ) cos2y−1 (θ)dθ. Then we compute

22

Page 24: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

that

∫Rnf(x)dnx = ∫

0∫

π

0⋯∫

π

0∫

0f(rv)rn−1 sinn−2 (φ1) sinn−3 (φ2) . . . sin (φn−2)dr dφ1 dφ2 . . .dφn−1

= (∫2π

0dφ1)(∫

π

0sinn−2 (φ2)dφ2) . . .(∫

π

0sin (φn−2)dφn−2)∫

0f(rv)rn−1 dr

= 2π (2∫π2

0sinn−2 (φ2)dφ2) . . .(2∫

π2

0sin (φn−2)dφn−2)∫

0f(rv)rn−1 dr

= 2πβ (n − 1

2,1

2)β (n − 2

2,1

2) . . . β (1,

1

2)∫

0f(rv)rn−1 dr

= 2πΓ (n−1

2)Γ ( 1

2)Γ (n−2

2)Γ ( 1

2)⋯Γ(1)Γ ( 1

2)

Γ (n2)Γ (n−1

2)⋯Γ ( 3

2) ∫

0f(rv)rn−1 dr

=2πΓ ( 1

2)n−2

Γ (n2) ∫

0f(rv)rn−1 dr

= 2πn2

Γ (n2) ∫

0f(rv)rn−1 dr

= Vol(Sn−1)∫∞

0f(rv)rn−1 dr.

Lemma 5.6. For any r ∈ Z, define fr = ure−u. Then for k ∈ N we have

∂k

∂ukfr =

k

∑j=0

(−1)j+k Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l

∂uk−lfr−j .

Proof. We proceed by induction. The case l = 0 is clearly true. For 0 < l < k, we have that

l

∑j=0

(−1)j+l Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l

∂uk−lfr−j =

l

∑j=0

(−1)j+l Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠((r − j) ∂

k−l−1

∂uk−l−1fr−j−1 −

∂k−l−1

∂uk−l−1fr−j)

=l

∑j=0

(−1)j+lΓ (r + 1)Γ (r − j)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l−1

∂uk−l−1fr−j−1

−l

∑j=0

(−1)j+l Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l−1

∂uk−l−1fr−j .

We now make the change of indices n = j + 1 on the first sum and relabel the second sum. This gives

l

∑j=0

(−1)j+l Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l

∂uk−lfr−j =

l+1

∑n=1

(−1)n+l+1 Γ (r + 1)Γ (r − n + 1)

⎛⎜⎝

l

n − 1

⎞⎟⎠∂k−l−1

∂uk−l−1fr−n

−l

∑n=0

(−1)n+l Γ (r + 1)Γ (r − n + 1)

⎛⎜⎝

l

n

⎞⎟⎠∂k−l−1

∂uk−l−1fr−n.

23

Page 25: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Figure 1: Domain of Integration

We now combine these into a single sum so that we may use the identity⎛⎜⎝

l + 1

n

⎞⎟⎠=⎛⎜⎝

l

n − 1

⎞⎟⎠+⎛⎜⎝

l

n

⎞⎟⎠

. This gives

l

∑j=0

(−1)j+l Γ (r + 1)Γ (r − j + 1)

⎛⎜⎝

l

j

⎞⎟⎠∂k−l

∂uk−lfr−j =

l

∑n=1

(−1)n+l+1 Γ (r + 1)Γ (r − n + 1)

⎛⎜⎝

⎛⎜⎝

l

n − 1

⎞⎟⎠+⎛⎜⎝

l

n

⎞⎟⎠

⎞⎟⎠∂k−l−1

∂uk−l−1fr−n

+ Γ (r + 1)Γ (r − l)

∂k−l−1

∂uk−l−1fr−l−1 +

∂k−l−1

∂uk−l−1fr

=l+1

∑n=0

(−1)n+l+1 Γ (r + 1)Γ (r − n + 1)

⎛⎜⎝

⎛⎜⎝

l

n − 1

⎞⎟⎠+⎛⎜⎝

l

n

⎞⎟⎠

⎞⎟⎠∂k−l−1

∂uk−l−1fr−n

=l+1

∑n=0

(−1)n+l+1 Γ (r + 1)Γ (r − n + 1)

⎛⎜⎝

l + 1

n

⎞⎟⎠∂k−l−1

∂uk−l−1fr−n.

Note that in the last equality, we have used the fact that for the terms n = 0 and n = l+1 there are only a single

term in the sum.

Example 5.7. We now attempt to compute

Trace (Mf (1 +D2APS,ε,R)

− s2 ) .

We will be computing the trace over the domain of the eigenvalues, so we will integrate over the region displayed

in Figure 1.

Note that the two lines have gradient c and −c respectively, where c = π

2(t2 − t1). Since there are no zero

eigenvalues, we introduce a lower limit ε for λ2 and a spatial cutoff R. The kernel for this operator is

Ks(x, y) = 2[ d+12

](2π)−d+12 f(x)gs(x − y),

24

Page 26: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

where g is defined as

gs(ξ) = (1 + ∥ξ∥2d − c2ξ2t )

− s2H (∥ξ∥2d − c

2ξ2t − ε2) .

Let cd = 2[ d+12

](2π)−d+12 , A = {(ξt, ξ1, . . . , ξd) ∈ Rd+1 ∶ ∥ξ∥2d − c2ξ2t > ε2} and B = {(r, p) ∈ R2 ∶ r2 − c2p2 > ε2}. Also

let ξ = (ξt, ξ1, . . . , ξd), x = (xt, x1, . . . , xd) and ∥ξ∥2d =d

∑k=1

ξ2k. Then we compute that

Trace (Mf (1 +D2APS,ε,R)

− s2 ) = ∫Rd+1

Ks(x,x)dx

= cd ∫Rd+1

Mf gs(0)dx

= cd ∫Rd+1

Mf(2π)−d+12 ∫

Rd+1(1 + ∥ξ∥2d − c

2ξ2t )− s2

dξ dx

= 2[ d+12

]

(2π)d+1 ∫Rd+1f(x)dx∫

A(1 + ∥ξ∥2d − c

2ξ2t )− s2

dξ.

We now carefully calculate the rightmost integral, using an analogue of Lemma 5.5, as

∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

dξ = 1

Γ ( s2) ∫

0us2−1 ∫

Ae−u(1+∥ξ∥

2d−ξ

2t ) dξ du

= 1

Γ ( s2) ∫

0us2−1∬

BVol (Sd−1) rd−1e−u(1+r

2−c2p2) dr dpdu.

We will need a change of coordinates. Let v = r− cp and w = r+ cp, so that v+w = 2r and w−v = 2cp. Calculate

the Jacobian as

J(r, p) = det⎛⎜⎝

∂v∂r

∂v∂p

∂w∂r

∂w∂p

⎞⎟⎠

= det⎛⎜⎝

1 c

−c 1

⎞⎟⎠

= 1 + c2.

Figure 5.7 demonstrates the coordinate transformation given above. Thus we have that

∬B

Vol (Sd−1) rd−1e−u(1+r2−c2p2) dr dp = 1

2 (1 + c2) ∫R

ε∫

w

ε2

w

(1

2(v +w))

d−1

e−uvw dv dw

+ 1

2 (1 + c2) ∫2R

R∫

2R−w

ε2

w

(1

2(v +w))

d−1

e−uvw dv dw

= 1

2d (1 + c2) ∫R

ε∫

w

ε2

w

d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠vkwd−1−ke−uvw dv dw (5.2)

+ 1

2d (1 + c2) ∫2R

R∫

2R−w

ε2

w

d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠vkwd−1−ke−uvw dv dw. (5.3)

We may now note that vke−uvw = (−1

w)k ∂k

∂uke−uvw and apply Lemma 5.6 to further compute the first integral

25

Page 27: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Figure 2: Coordinate transformation applied to the domain of integration

5.2 as (for 2k ≠ d − 1)

∫R

ε∫

w

ε2

w

wd−1−kvke−uvw dv dw = ∫R

ε∫

w

ε2

w

(−1)kwd−1−2k ∂k

∂uke−uvw dv dw

= (−1)k ∂k

∂uk∫

R

εwd−2−2k ( 1

u(e−uε

2

− e−uw2

))dw

= (−1)k ∂k

∂uk( 1

ue−uε

2

)( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

− (−1)k ∂k

∂uk∫

R

εwd−2−2k

e−uw2

udw.

To work with these integrals, we recall that they were inside an integral over u. Then for s large and using

integration by parts we find that

∫∞

0us2−1e−u ∫

R

ε∫

w

ε2

w

wd−1−kvke−uvw dv dw du = ∫∞

0(−1)ku

s2−1e−u

∂k

∂uk⎛⎝

e−uε2

u

⎞⎠

du( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

− ∫∞

0(−1)ku

s2−1e−u

∂k

∂uk∫

R

εwd−2−2k

e−uw2

udw du

=k

∑j=0

cj,k,s ∫∞

0us2−j−2e−ue−uε

2

( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)du

−k

∑j=0

cj,k,s ∫R

ε∫

0us2−j−2e−ue−uw

2

wd−2−2k dudw

=k

∑j=0

cj,k,s (1 + ε2)−s2+j+1 ( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

−k

∑j=0

cj,k,s ∫R

ε(1 +w2)−

s2+j+1wd−2−2k dw

where cj,k,s = (−1)j+kΓ ( s

2)

Γ ( s2− j)

⎛⎜⎝

k

j

⎞⎟⎠

. Using integration by parts we see that this is much more computable when

d − 2 − 2k is odd than even.

If 2k = d − 1 then the same calculation applies, however every term of the form ( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

26

Page 28: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

becomes ln(Rε), so that

∫∞

0us2−1e−u ∫

R

ε∫

w

ε2

w

wd−1−kvke−uvw dv dw du =k

∑j=0

cj,k,s (1 + ε2)−s2+j+1 ln(R

ε)

−k

∑j=0

cj,k,s ∫R

ε(1 +w2)−

s2+j+1wd−2−2k dw

We compute this last integral as

∫R

ε(1 +w2)−

s2+j+1wd−2−2k dw = ∫

R

ε

1

Γ ( s2− j − 1) ∫

0us2−j−2e−u(1+w

2)wd−2−2k dw du

= 1

Γ ( s2− j − 1) ∫

0us2−j−2e−u(−1)

d−2−2k2

∂d−2−2k

2

∂u2−d−2k

2∫

R

εwP (d−2−2k)e−u(1+w

2) dw du

= 1

Γ ( s2− j − 1) ∫

0

∂d−2−2k

2

∂ud−2−2k

2

(us2−j−1e−u)md,k(u)du

= 1

Γ ( s2− j − 1)

[ d−2−2k2

]

∑p=0

(−1)[d−2−2k

2]+p

Γ ( s2− j − 1)

Γ ( s2− j − 1 + p)

⎛⎜⎝

[d−2−2k2

]

p

⎞⎟⎠Md,k,s,p,

where P is the parity, giving 1 if odd and 0 if even,

md,k(u) =

⎧⎪⎪⎪⎪⎨⎪⎪⎪⎪⎩

√π2(erf (

√uR) − erf (

√uε)) if d − 2 − 2k is even,

12(e−uR

2

− e−uε2

) if d − 2 − 2k is odd.

and

Md,k,s,p = ∫∞

0us2−j−1−p−1e−umd,k(u)du.

We may now also compute in a similar manner the integral 5.3 as

∫2R

R∫

2R−w

ε2

w

wd−1−kvke−uvw dv dw = (−1)k ∫2R

R∫

2R−w

ε2

w

wd−1−2k∂k

∂uk(e−uvw)dv dw

= (−1)k ∂k

∂uk∫

2R

R

wd−1−2k

uw(e−uε

2

− e−uw(2R−w))dw

= (−1)k ∂k

∂uk⎛⎝

e−uε2

u

⎞⎠((2R)d−1−2k

d − 1 − 2k− Rd−1−2k

d − 1 − 2k)

− (−1)k ∫2R

R

wd−2−k

ue−uw(2R−w) dw

= (−1)k ∂k

∂uk⎛⎝

e−uε2

u

⎞⎠((2R)d−1−2k

d − 1 − 2k− Rd−1−2k

d − 1 − 2k) − Id,k.

Note that the case 2k = d−1 gives rise to a logarithm term, as in the previous integral. For further computation,

27

Page 29: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

we again bring back the integral over u and apply integration by parts. This gives

∫∞

0us2−1e−uId,k du = (−1)k ∫

0us2−1e−u

∂k

∂uk∫

2R

R

wd−2−2k

ue−uw(2R−2) dw du

= ∫∞

0

∂k

∂uk(u

s2−1e−u)∫

2R

R

wd−2−2k

ue−uw(2R−w) dw du

=k

∑j=0

(−1)j+kΓ ( s

2)

Γ ( s2− j)

⎛⎜⎝

k

j

⎞⎟⎠∫

0us2−2−je−u(1+w(2R−w)) ∫

2R

Rwd−2−2k dw du

=k

∑j=0

(−1)j+kΓ ( s

2)

s2− j − 1

⎛⎜⎝

k

j

⎞⎟⎠∫

2R

R(1 +w(2R −w))−

s2+1+j wd−2−2k dw.

Combining these calculations together, we see that

∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

dξ =d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+kΓ ( s

2)

s2− j − 1

⎛⎜⎝

k

j

⎞⎟⎠(1 + ε2)−

s2+j+1 ( Rd−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

−d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+k⎛⎜⎝

k

j

⎞⎟⎠

Γ ( s2)

s2− j − 1

∫R

ε(1 +w2)−

s2+j+1wd−2−2k dw

+d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+k⎛⎜⎝

k

j

⎞⎟⎠

Γ ( s2)

s2− j − 1

(1 + ε2)−s2+j+1 ((2R)d−1−2k

d − 1 − 2k− Rd−1−2k

d − 1 − 2k)

−d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+kΓ ( s

2)

s2− j − 1

⎛⎜⎝

k

j

⎞⎟⎠∫

2R

R(1 +w(2R −w))−

s2+j+1wd−2−2k dw

=d−1

∑k=0

k

∑j=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

⎛⎜⎝

k

j

⎞⎟⎠(−1)j+k

Γ ( s2)

s2− j − 1

(1 + ε2)−s2+j+1 ((2R)d−1−2k

d − 1 − 2k− εd−1−2k

d − 1 − 2k)

−d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+k⎛⎜⎝

k

j

⎞⎟⎠

Γ ( s2)

s2− j − 1

∫R

ε(1 +w2)−

s2+j+1wd−2−2k dw

−d−1

∑k=0

⎛⎜⎝

d − 1

k

⎞⎟⎠

k

∑j=0

(−1)j+kΓ ( s

2)

s2− j − 1

⎛⎜⎝

k

j

⎞⎟⎠∫

2R

R(1 +w(2R −w))−

s2+j+1wd−2−2k dw

and so

Trace (Mf (1 +D2)−s2 ) =

2[ d+12

]−dVol (Sd−1)(2π)d+1 (1 + c2) ∫Rd+1

f(x)dx∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

dξ.

We will now compute this trace for some low dimensional examples from d = 1 to d = 3. First we consider d

odd since the integrals are simpler.

Example 5.8 (Dimension d + 1 = 2). We will consider the expression for I(s) = ∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

computed in example 5.7. Note that there is a simple pole at s = 2. We compute the residue at this pole to

determine its behaviour.

Ress=2(I(s)) = ln(2R

ε) − (R − ε) − (2R −R)

= ln(2R

ε) − 2R + ε.

28

Page 30: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

Example 5.9 (Dimension d + 1 = 4). We will consider the expression for I(s) = ∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

computed in example 5.7. Note that there are simple poles at s = 2, s = 4 and s = 6. We compute the residue

at these poles to determine their behaviour.

For the pole at s = 2, we only need to consider terms with j = 0. Hence we have

Ress=2(I(s)) = −2R2 + ε2

2− ln(2R

ε) + (2R)−1

2− ε

−1

2− (2R)−2 + ε−2.

For the pole at s = 4, we only need to consider terms with j = 1. Thus we have

Ress=4(I(s)) =⎛⎜⎝

2

1

⎞⎟⎠

⎛⎜⎝

1

1

⎞⎟⎠

Γ (2) (1 + ε2)0 ln(Rε) +

⎛⎜⎝

2

2

⎞⎟⎠

⎛⎜⎝

2

1

⎞⎟⎠(−1)Γ (2) (1 + ε2)0 ((2R)−2

−2− ε

−2

−2)

−⎛⎜⎝

2

1

⎞⎟⎠

⎛⎜⎝

1

1

⎞⎟⎠

Γ (2) (ln(Rε) + ln(2R

R)) + 2

⎛⎜⎝

2

2

⎞⎟⎠

⎛⎜⎝

2

1

⎞⎟⎠

Γ (2)∫R

εw−3 dw

= 4 ln(Rε) − 4 ln(R

ε) − 4 ln (2) + 4 (2R−2 − ε−2) − (R−4 − ε−4)

= −4 ln (2) + 8R−2 − 2R−4 − 4ε−2 + 2ε−4.

For the pole at s = 6, we only need to consider terms with j = 2. We compute this as

Ress=6(I(s)) =⎛⎜⎝

2

2

⎞⎟⎠

⎛⎜⎝

2

2

⎞⎟⎠(−1)2+2Γ (4)((2R)−1

−1− ε

−1

−1− ∫

R

εw−2 dw − ∫

2R

Rw−2 dw)

= 6 (−(2R)−1 + ε−1 − ε−1 + (2R)−1)

= 0.

We now compute an odd example with d + 1 = 3.

Example 5.10 (Dimension d + 1 = 3). We will consider the expression for I(s) = ∫A(1 + ∥ξ∥2d − c

2ξ2t )− s2

computed in example 5.7. Note that there is a simple pole at s = 2 and at s = 4. We compute the residue at

these poles to determine their behaviour.

For s = 2, we only need to consider terms with j = 0. Calculate the residue as

Ress=2(I(s)) =⎛⎜⎝

1

1

⎞⎟⎠

⎛⎜⎝

1

0

⎞⎟⎠

Γ (1) [ε−1 − (2R)−1 − ∫R

εw−2 dw − ∫

2R

Rw−2 dw]

+⎛⎜⎝

1

0

⎞⎟⎠

⎛⎜⎝

0

0

⎞⎟⎠[2R − ε − ∫

R

εw dw − ∫

2R

Rw dw]

= ε−1 − (2R)−1 − ε−1 + (2R)−1 + 2R − ε − 1

2(2R)2 + 1

2ε2

= 2R − ε − 1

2(2R)2 + 1

2ε2.

29

Page 31: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

For s = 4, we need only consider terms with j = 1. This residue is calculated as

Ress=4(I(s)) = −⎛⎜⎝

2

2

⎞⎟⎠

⎛⎜⎝

2

1

⎞⎟⎠[ε

−2 − (2R)−2

2− ∫

R

εw−3 dw − ∫

2R

Rw−3 dw]

+⎛⎜⎝

2

1

⎞⎟⎠

⎛⎜⎝

1

1

⎞⎟⎠[ln(2R

ε) − ∫

R

εw−1 dw − ∫

2R

Rw−1 dw]

= 0.

It should be noted that in all of the above examples, taking the residue made the integral computations

significantly easier, removing the distinction between even and odd. We would also like to take the limit as

R → ∞, which is not expected to commute with the operation of taking a residue, hence these results can be

expected to have lost information about the problem.

Example 5.11. Another possible candidate for our trace is Trace (MfD2 (1 +D∗D)−

s2 ). Let ξ = (ξt, ξ1, . . . , ξd),

x = (xt, x1, . . . , xd) and y = (yt, y1, . . . , yd) then compute, using analogous results to those found in [9, Corollary

14], and Example 5.7 that

Trace (MfD2(1 +D∗D)−

s2 ) = 2[ d+1

2](2π)−(d+1) ∫

Rd+1∫Af(x)

⎛⎝− ∂2

∂x2t+

d

∑j=1

∂2

∂x2j

⎞⎠[e−i(x−y)⋅ξ (1 + ∥ξ∥2)

− s2 ]x=y

dξ dx

= cd ∫Rd+1∫A(f(x) (ξ2t − ∥ξ∥2d) (1 + ξ2t + ∥ξ∥2d)

− s2e−i(x−y)⋅ξ)

x=ydξ dx

= cd ∫A((1 + ξ2t + ∥ξ∥2d)

− s2 (2ξ2t + 1) − (1 + ξ2t + ∥ξ∥2d)− s2+1)dξ∫

Rd+1f(x)dx,

where cd = 2[ d+12

](2π)−(d+1) and A = {(ξt, ξ1, . . . , ξd) ∈ Rd+1 ∶ ∥ξ∥2d − c2ξ2t > ε2}. We will calculate these two

integrals separately. Let B = {(r, p) ∈ R2 ∶ r2 − c2p2 > ε2} and compute that

∫A(1 + ξ2t + ∥ξ∥2d)

− s2 (2ξ2t + 1)dξ = 1

Γ ( s2) ∫

0∫Aus2−1 (2ξ2t + 1) e−u(1+ξ

2t+∥ξ∥

2d) dξ du

= Vol (Sd−1)∫∞

0∬

Bus2−1 (2p2 + 1) rd−1e−u(1+r

2+p2) dpdr du.

We will evaluate the integrals separately and recombine. First, consider the r integral and note that the region

of integration is D = (−∞,−√ε2 + c2p2) ∪ (

√ε2 + c2p2,∞). Then we have that

Id = ∫Drd−1e−ur

2

dr

= ∫D−r

d−2

2u

d

dr(e−ur

2

)dr

=⎛⎝−r

d−2e−ur2

2u

⎞⎠∣∂D + d − 2

2u∫Drd−2e−ur

2

dr

= e−u(ε2+c2p2) ((

√ε2 + c2p2)

d−2− (−

√ε2 + c2p2)

d−2) + d − 2

2uId−3.

30

Page 32: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

So we calculate I2 as

I2 = ∫Dre−ur

2

dr

= −1

2(e−ur

2

) ∣∂D

= 0

and I1 as

I1 = ∫D

e−ur2

dr

=√π

uerfc (

√ε2 + c2p2) .

This integral appears quite difficult to evaluate and so we will not attempt to continue the computation.

6 Conclusion

In this paper we developed the theory of differential operators on manifolds and clifford algebras in order to

define and work with Dirac operators, of particular interest in physics. The main result of this report is that a

trace can be computed for the Dirac operator on a Lorentzian cylinder. Further work could include analysing

the trace in the limit as R →∞ before taking the residues and continuing with the integral computations found

in Example 5.11. A further extension would be to consider a compact manifold with more interesting topology,

such as a torus.

31

Page 33: Dirac Operators on Manifolds and Applications to Physics · 1 Introduction This report develops the prerequisite material for studying Dirac operators on Lorentzian manifolds and

References

[1] Christian Bar, Nicolas Ginoux, and Frank Pfaffle. Wave Equations on Lorentzian Manifolds and Quanti-

zation. European Mathematical Society, 2007.

[2] Christian Bar and Alexander Strohmaier. An index theorem for lorentzian manifolds with compact spacelike

cauchy boundary. American Journal of Mathematics, 2015.

[3] Alain Connes. Gravity coupled with matter and the foundation of non commutative geometry. Communi-

cations in Mathematical Physics, (1):155–156, 1996.

[4] R.W.R. Darling. Differential Forms and Connections. Cambridge University Press, 1994.

[5] Nigel Higson and John Roe. Analytic K-Homology. Oxford Mathematical Monographs, OUP, 2000.

[6] Jurgen Jost. Riemannian Geometry and Geometric Analysis. Springer, 2005.

[7] H. Blaine Lawson and Marie-Louise Michelsohn. Spin Geometry. Princeton University Press, 1989.

[8] Barrett O’Neill. Semi-Riemannian Geometry with Applications to Relativity. Academic Press, 1983.

[9] Adam Rennie. Summability for nonunital spectral triples. K-Theory, (31):71–100, 2004.

[10] R. E. Showalter. Monotone Operators on Banach Space and Nonlinear Partial Differential Equations.

American Mathematical Society, 1997.

[11] Wolfgang Yourgrau and Stanley Mandelstam. Variational Principles in Dynamics and Quantum Theory.

General Publishing Company, 1968.

32