139
CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A Dissertation Submitted to the Graduate School of the University of Notre Dame in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy by D. Corbett Redden, B.A., M.S. Stephan Stolz, Director Graduate Program in Mathematics Notre Dame, Indiana July 2006

CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING

STRUCTURES

A Dissertation

Submitted to the Graduate School

of the University of Notre Dame

in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

by

D. Corbett Redden, B.A., M.S.

Stephan Stolz, Director

Graduate Program in Mathematics

Notre Dame, Indiana

July 2006

Page 2: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

c© Copyright by

D. Corbett Redden

2006

All Rights Reserved

Page 3: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING

STRUCTURES

Abstract

by

D. Corbett Redden

In this thesis, we look at principal Spin(n)-bundles P → M whose Pontryagin

class p1

2(P ) = 0 ∈ H4(M ; R). We say that such a bundle admits a string structure,

and a choice of string structure is given by particular elements in H3(P ; Z). This

provides an analogue to the idea of a spin structure, where the topological group

String(n) is, up to homotopy, the unique 3-connected cover of Spin(n) (n ≥ 5).

Choosing a Riemannian metric on the base M and a connection on P determines

a 1-parameter family of metrics on P . We prove that in a scaling limit known as

the adiabatic limit, the harmonic representative of a string structure is equal to

the Chern–Simons 3-form on P minus a 3-form on M , denoted H ∈ Ω3(M). This

3-form H is closely related to the Chern–Simons form, and can be thought of as a

reduction of the Chern–Simons form on P to a form on M . The exterior derivative

of H is the p1

2-form; the integral of H on any 3-cycle is equal, modulo Z, to the

integral of the Chern–Simons 3-form pulled back via a global section on the same 3-

cycle. Finally, we note that for the Spin(n)-frame bundle Spin(M) of a Riemannian

spin manifold, this 3-form H determines a canonical metric connection on M . This

canonical metric connection depends on both the metric and string structure, and

its torsion is determined by H.

Page 4: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

To Mom and Dad.

ii

Page 5: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CONTENTS

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

CHAPTER 1: INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Original motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.3 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

CHAPTER 2: PRELIMINARIES . . . . . . . . . . . . . . . . . . . . . . . . . 172.1 Lie groups and torsors . . . . . . . . . . . . . . . . . . . . . . . . . . 172.2 Right invariant vector fields and forms on a G-torsor . . . . . . . . . 222.3 Connections on principal G-bundles . . . . . . . . . . . . . . . . . . . 252.4 Hodge theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312.5 Chern–Weil and Chern–Simons Forms . . . . . . . . . . . . . . . . . . 36

CHAPTER 3: RIGHT-INVARIANT FORMS ON A PRINCIPAL BUNDLE . 413.1 Isomorphic bi-graded cochain complex . . . . . . . . . . . . . . . . . 413.2 The adjoint d∗ in the bi-graded complex . . . . . . . . . . . . . . . . 523.3 Rescaled derivatives dδ and d∗δ . . . . . . . . . . . . . . . . . . . . . . 55

CHAPTER 4: ADIABATIC SPECTRAL SEQUENCE . . . . . . . . . . . . . 594.1 Leray–Serre spectral sequence . . . . . . . . . . . . . . . . . . . . . . 604.2 Adiabatic spectral sequence . . . . . . . . . . . . . . . . . . . . . . . 614.3 Relation to harmonic forms . . . . . . . . . . . . . . . . . . . . . . . 68

CHAPTER 5: HARMONIC FORMS ON A PRINCIPAL BUNDLE IN THEADIABATIC LIMIT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745.1 1-forms when G = U(n) . . . . . . . . . . . . . . . . . . . . . . . . . 755.2 1- and 2-forms when G is semisimple . . . . . . . . . . . . . . . . . . 835.3 3-forms when G is simple . . . . . . . . . . . . . . . . . . . . . . . . . 88

iii

Page 6: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 6: CANONICAL FORMS ASSOCIATED TO LIFTS OF THESTRUCTURE GROUP . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1006.1 Generalities about G′ structures . . . . . . . . . . . . . . . . . . . . . 1006.2 U structures on U(n)-bundles . . . . . . . . . . . . . . . . . . . . . . 1096.3 Spinc structures on SO(n)-bundles . . . . . . . . . . . . . . . . . . . 1116.4 String structures on Spin(n)-bundles . . . . . . . . . . . . . . . . . . 1136.5 Canonical forms on base from G′ structures . . . . . . . . . . . . . . 117

CHAPTER 7: CANONICAL METRIC CONNECTIONS ASSOCIATED TOSTRING STRUCTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

iv

Page 7: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

ACKNOWLEDGMENTS

First, I would like to thank my advisor, Stephan Stolz. In addition to his valuable

insight and interest in a wide variety of problems, his patience, availability, and

enthusiasm have created an ideal environment in which to develop. I would also

like to thank Brian Hall, Liviu Nicolaescu, and Bruce Williams. Not only did they

provide helpful comments as readers for this thesis, but I have also learned much

from courses they taught and discussions with them over the past five years. I

am also grateful to the math faculty and graduate students at Notre Dame for the

great mathematical environment they create. In particular, I want to thank Florin

Dumitrescu and Elke Markert for the innumerable discussions.

Thanks also to all the friends at Notre Dame who have made the past five years

so much fun. There are too many to list, so I will not bother trying. Finally, thanks

to my parents, Mike and Elaine Redden, and my brother Collin for the continued

love and support over all the years.

v

Page 8: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 1

INTRODUCTION

1.1 Summary

The Pontryagin class p1

2(M) ∈ H4(M ; Z), defined on a closed compact spin

manifold M , is a common obstruction to defining (even formally) certain operators

or constructions in quantum field theory. For example, the spinor bundle on the

loop space LM exists if p1

2(M) = 0, but not in general [CP]. Also, important

2-dimensional supersymmetric sigma models associated to a Riemannian manifold

with spin structure don’t even make formal sense unless p1

2(M) = 0 [AS1].

The definition of a string structure is not a universal one, but in this thesis, a

spin manifold M admits a string structure if and only if p1

2(M) = 0. Given a spin

structure on M , let Spin(M)π→ M be the Spin(n)-frame bundle (where n ≥ 5).

A string structure is then defined as a cohomology class S ∈ H3(Spin(M); Z) that

restricts to the (standard) generator of H3(Spin(n); Z) ∼= Z on the fibers. The

primary goal of this thesis is the following construction. Given a Riemannian spin

manifold (M, g), there is a natural 1-parameter family of metrics gδ on the bundle

Spin(M), and the limit as δ → 0 is referred to as the adiabatic limit. Given a

string structure S ∈ H3(Spin(M); Z), we take the harmonic representative in the

adiabatic limit of (the real cohomology class given by) S. By Theorem 6.4.4, this

form is equal to the Chern–Simons 3-form α(Θ) minus a 3-form Hg,S pulled back

1

Page 9: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

from M ; i.e.

limδ→0

[S]gδ= α(Θ)− π∗Hg,S ∈ Ω3(P ).

The form Hg,S ∈ Ω3(M) is closely related to the Chern–Simons 3-form on P , and

can be thought of as a reduction of the Chern–Simons 3-form to a form on M . In

particular, dHg,S = p1

2(M, g), and the harmonic component of Hg,S, once reduced

to H3(M ; R/Z), equals the 3-dimensional Chern–Simons cohomology class. The

forms α(Θ)−π∗Hg,S and Hg,S arise in both the construction of the spinor bundle on

LM and making formal sense of 2-dimensional supersymmetric sigma models with

target space M . In future work, the author hopes to formalize the link between

these constructions and the results in this thesis.

Given a choice of metric and string structure on M , we have the canonical 3-

form Hg,S, and we use this to construct a canonical 1-parameter family of metric

connections ∇εg,S on TM . These connections are characterized their torsion, which

is given by

〈T εg,S(X, Y ), Z〉g =1

εHg,S(X, Y, Z).

The limit as ε → ∞ gives the Levi-Civita connection, and the limit as ε → 0 is

not a connection for Hg,S 6= 0. Connections with torsion have become important

in string theory. Though there are similarities between this family of connections

and other connections appearing in the literature, the precise relationships are not

yet known by the author. The hope is that geometric properties of this family of

connections is related to invariants coming from conformal field theories and elliptic

cohomology. This motivated the development of these connections, but there is

not yet any mathematical link. We will go ahead and briefly describe the possible

relationship, but it should be noted that this is only one possible use of the results

in this thesis.

2

Page 10: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

1.2 Original motivation

In [Sto], Stolz conjectures that if a manifold M admits both a metric of positive

Ricci curvature and a string structure, then a topological invariant called the Witten

genus vanishes; i.e. φW (M) = 0. While this remains a valid conjecture, the naive

guess that positive Ricci curvature could imply the vanishing of the refined Witten

genus, which lives in the cohomology theory tmf , is simply not true. The hope is

that some variant of this statement holds true when considering the curvatures of

the canonical 1-parameter family of connections depending on both the metric and

the string structure.

There is a simpler but analogous situation. If a spin manifold M admits a metric

of positive scalar curvature s > 0, then A(M) = 0 [Lic]. This can be seen by the the

Atiyah-Singer Index Theorem and the Bochner-Weizenbock-Lichnerowicz formula

6D2 = ∇∗∇ +s

4[LM]. If s > 0, then 6D2 is a strictly positive operator, and hence

its kernel and A(M) are both 0. There is also a refinement of A given by the spin-

orientation of KO-theory. This induces a map α : ΩSpinn (pt) → KO−n(pt), which

depends on the choice of spin structure, such that the following diagram commutes:

KO−n(pt)

ΩSpin

n (pt)bA //

α88ppppppppppp

Z.

For a spin structure d, the element α[M,d] ∈ KO−n(pt), sometimes called the

Atiyah α-invariant, can be thought of as the Clifford-linear index of 6DM . The

above construction usually appears in family index theorems, but it also contains

interesting information for a single manifold due to the torsion in KO−∗(pt). By a

theorem of Hitchin, one can use a similar proof to show that if M admits a positive

scalar curvature metric, then α[M,d] = 0 ∈ KO−n(pt) for any spin structure d [Hit].

3

Page 11: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This result has a nice interpretation within the context of quantum field theory.

Stolz and Teichner have shown that KO−n ' EFTn, i.e. that the (−n)-th space

in the KO-theory spectrum is homotopy equivalent to the space of supersymmetric

1-dimensional Euclidean Field Theories of degree n [ST]. Roughly speaking, one can

interpret the map α as, up to homotopy, quantizing fermionic particles moving in a

target spin manifold M . The integer A(M) is the partition function of the resulting

field theory. Hitchin’s theorem implies that if Mn admits a positive scalar curvature

metric, the 1-dimensional field theory associated to Mn is qualitatively the same as

the trivial field theory of degree n. From the perspective of Stolz and Teichner, the

relationship between KO-theory and 1-dimensional Euclidean field theories should

be analogous to the relationship between 2-dimensional conformal field theories and

elliptic cohomology.

As mentioned above, Stolz conjectures that the following analogous situation

should hold for string manifolds: if M admits a string-structure and a metric of

positive Ricci curvature, then the Witten genus φW (M) = 0 [Sto]. His heuristic

argument comes from interpreting the Witten genus as the S1-equivariant index

of the Dirac operator on the free loop space LM (i.e. φW (M) = indexS1(6DLM))

[Wit2][Wit3]. Also, there should be some Weizenbock-type formula involvingRic(M)

such that if Ric(M) > 0, then Ker(6DLM) = 0. This line of argument is far from

rigorous since 6DLM and the scalar curvature on LM are not well-defined mathe-

matical objects. However, the conjecture holds true for homogeneous spaces and

complete intersections. Currently, there are no known examples of simply connected

manifolds which admit metrics of positive scalar curvature, but not metrics of pos-

itive Ricci curvature. If the conjecture is true, it would provide examples of such

manifolds.

Just as the cohomology theoryKO is the home for the A-genus of a spin manifold

4

Page 12: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

(or a family of spin manifolds), there is a cohomology theory tmf (topological mod-

ular forms) that is the natural home for the Witten genus φW of a string manifold

([Hop], [AHS]). In other words, there is a map σ such that the following diagram

commutes:

tmf−n(pt)

ΩString

n (pt)φW //

σ77oooooooooooMFn

The map tmf−∗(pt)→MF∗ is a rational isomorphism, whereMFn denotes modular

forms of weight n2. There are several attempts to geometrically define tmf (or

other elliptic cohomology theories) in terms of 2-dimensional conformal field theories

(including [Seg1], [Seg2], and [ST]). Under this view, σ(M) is (up to homotopy)

the super-symmetric non-linear sigma model for a target space M . The partition

function would be φW (M).

It is important to note that both A(M) and φW (M) are topologically defined

and thus independent of spin or string structure. However, the maps α and σ into

the appropriate cohomology theories do depend on such choices. Now, one might

hope that the analog of Hitchin’s theorem holds. In other words, is it possible that

if Mn admits a metric of positive Ricci curvature, then σ[Mn, S] = 0 ∈ tmf−n(pt)

for all string structures S? The answer is no, and counterexamples can be found by

looking at Lie groups.

Assume G is a compact, semi-simple Lie group. G admits a bi-invariant metric

which has positive Ricci (and sectional) curvature. G also has two canonical string

structures defined by the left and right invariant framing of TG, which we will denote

L and R. The image of these under σ has been computed in many examples and

are often not 0 [Hop], hence the above question cannot have an affirmative answer.

The easiest case to see is SU(2) ∼= S3. In addition to the string structures coming

5

Page 13: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

from left and right invariant framing, there is the trivial string structure given by

S3 = ∂D4. We then have the following:

ΩString3

σ−→ tmf−3 ∼= Z/24

[SU(2),L] 7−→ −1

24

[SU(2), ∂D4] 7−→ 0

[SU(2),R] 7−→ 1

24

The above follows from the fact that the maps π3S0 → π3MString → π3tmf are

all isomorphisms [Hop]. This reduces to a framed bordism calculation, and using

the Adams e-invariant, one can show that the different framings on SU(2) give the

above framed bordism classes [AS2].

Using Proposition 6.5.1 and the H3(M ; Z)-equivariance of our construction, we

see that for the bounding string structure ∂D4, the 3-form given by our construction

is

Hg,∂D4 = 0 ∈ Ω3(SU(2)).

Therefore, the resulting 1-parameter family of connections ∇εg,∂D4is always the

Levi-Civita connection, and hence always has positive curvature. However, the

1-parameter family of connections ∇εg,L, ∇εg,R have negative Ricci curvature for

ε small. From the description of the Ricci curvature in (7.0.1), we see that in

general, if Hg,S 6= 0, then Ric∇εg,S → −∞ as ε → 0. The most obvious question

is: if Ric∇εg,S > 0 for all ε, does this imply that σ(M, S) = 0 ∈ tmf−n(pt)?

Equivalently, if (M, g) has positive Ricci curvature and Hg,S = 0, does this imply

that σ(M, S) = 0?

6

Page 14: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

1.3 Outline of the thesis

We now proceed to give a more in-depth outline of the actual thesis. Chapters 2,

3, and 4 all build up to the calculation in Chapter 5 of the 3-dimensional harmonic

forms on a principal G-bundle (G simple) in the adiabatic limit. This calculation

leads to the canonical 3-forms on M associated to a metric and string structure (and

connection). Finally, the canonical 3-forms give rise to the canonical connections in

Chapter 7.

All manifolds are considered to be closed and compact. Let Pπ→ M be a

principal G-bundle, for G a compact Lie group. We begin in Chapter 2 by reviewing

the theory of connections on principal bundles, the basics of Hodge theory, as well

as the Chern–Simons and Chern–Weil forms we use. On any principal bundle, there

is a canonical distribution of “vertical” vectors given by

T V Pdef= Ker(π∗) ⊂ TP.

A choice of connection is equivalent to the choice of an equivariant distribution of

“horizontal” vectors. Letting Θ denote the connection 1-form on P , this is written

as

THPdef= Ker Θ ⊂ TP.

Therefore, a connection induces a bi-grading on the space of differential forms,

denoted

Ωi,j(P )def= C∞(P,ΛiTHP ∗ ⊗ ΛjT V P ∗).

This bi-grading is extremely helpful in later calculations. A connection also induces

Chern–Weil forms on M and Chern–Simons forms on P . In particular, for G simple,

the Chern–Simons 3-form is denoted α(Θ) ∈ Ω3(P ), and its derivative is the pullback

of the Chern–Weil form 〈Ω ∧Ω〉 ∈ Ω4(M) (where 〈·, ·〉 is a suitably normalized Ad-

7

Page 15: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

invariant metric on g, and Ω is the curvature 2-form).

By definition, G acts freely from the right on P . If the metric on P is G-

equivariant, then the harmonic forms will be G-invariant. Therefore, in Chapter 3,

we give a useful description of the subcomplex of right-invariant differential forms on

P . In addition to aiding calculations, the author finds this description conceptually

helpful; it relates the exterior derivative on P to the Lie algebra derivative, the

connection Θ, and the curvature Ω.

The description is given by an isomorphism of cochain complexes. To see this,

first note that there is a canonical isomorphism of G-equivariant vector bundles

T V P ∼= P × g = π∗gP ,

where gP = P ×Ad g is the adjoint vector bundle over M . Given a connection Θ,

THP = π∗TM,

and thus we have a canonical isomorphism

Ωi,j(P )G ≈←→ Ωi(M ; Λjg∗P ).

The exterior derivative on Ωi,j(P ) decomposes with respect to the bi-grading, and

d = d0,1 + d1,0 + d2,−1

where da,b : Ωi,j(P ) → Ωi+a,j+b(P ). When restricted to right-invariant forms, the

derivations da,b are given by familiar maps on the isomorphic complex Ωi(M ; Λjg∗P ).

The derivative of right-invariant forms on the fibers induces the vector bundle ho-

momorphism

Λjg∗Pdg−→ Λj+1g∗P .

8

Page 16: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This gives the map

d0,1 = (−1)idg : Ωi(M ; Λjg∗P )→ Ωi(M ; Λj+1g∗P ).

The principal bundle connection Θ induces a vector bundle connection ∇ on each

associated vector bundle Λjg∗P , and

d1,0 = d∇ : Ωi(M ; Λjg∗P )→ Ωi+1(M ; Λjg∗P ).

Finally, the curvature of a connection is an element Ω ∈ Ω2(M ; gP ), and d2,−1 is

given by contracting along this vector-valued 2-form; this is denoted

d2,−1 = (−1)iιΩ : Ωi(M ; Λjg∗P )→ Ωi+2(M ; Λj−1g∗P ).

In particular, d2,−1 = 0 if and only if the connection is flat.

To make sense of Hodge cohomology, P must be a Riemannian manifold. Given

a Riemannian metric g on M and a connection on P , we see a natural family of

metrics on P . First, choose a bi-invariant metric gG on G, or equivalently, choose an

Ad-invariant metric on g. Such metrics exist because G is compact, and all theorems

in Chapter 5 are independent of this choice. Defining the vertical and horizontal

spaces to be perpendicular then gives the metric gP , written

gPdef= π∗(g ⊕ gG)

under the decomposition TP ∼= π∗(TM ⊕ gP ).

The main result of this thesis follows from the 3-dimensional Hodge cohomology

of P . However, these forms only have a nice description under the scaling limit

given by shrinking the fibers relative to the base, or enlarging the base relative to

the fibers. Introducting this scaling factor δ, define

gδdef= π∗

(δ−2g ⊕ gG

).

9

Page 17: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

The limit δ → 0 is referred to as the adiabatic limit, and was introduced by Witten

in [Wit1]. It has produced a number of interesting results, including [BF]. Though

g0 is not a metric, the work of [MM] and [For] gives a smooth extension of Ker ∆gδ

to δ = 0. The goal of Chapter 5 is to describe

limδ→0

Ker ∆3gδ⊂ Ω3(P ).

Note that for any δ > 0, gδ is a right-invariant Riemannian metric on P , and is given

by the pullback of an inner product on Ωi(M ; Λjg∗P ). The adjoint d∗gδ on Ω∗(P ),

when restricted to right-invariant forms, then decomposes in terms of the adjoints

of dg, d∇, and ιΩ.

Chapter 4 summarizes results concerning a Hodge-theoretic version of the Leray–

Serre spectral sequence. This spectral sequence first appeared in [MM] and was later

made explicit in [For]. A key role is played by the isometry

ρδ : (Ωi,j(P ), gδ) −→ (Ωi,j(P ), gP )

φ 7−→ δiφ.

In addition to fixing the inner product space, this also induces factors of δ which

give the structure of a spectral sequence. Conjugating by the isometry ρδ gives a

rescaled derivative dδ, a rescaled coderivative d∗δ , and a rescaled Laplacian Lgδ.

dδdef= ρδdρ

−1δ = d0,1 + δd1,0 + δ2d2,−1,

d∗δdef= ρδd

∗gδρ−1δ = d0,1 + δd1,0 + δ2d2,−1,

Lgδ

def= ρδ∆gδ

ρ−1δ = dδd

∗δ + d∗δdδ.

10

Page 18: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

When restricted to right-invariant forms, dδ and d∗δ take the form

dδ = ±dg + δd∇ ± δ2ιΩ,

d∗δ = ±d∗g + δd∗∇ ± δ2ι∗Ω.

Treating δ as a formal variable, there is a filtration of forms on P given by

Ei,jK =ω ∈ Ωi,j(P )|∃ ω1, . . . , ωl with

dδ(ω + δω1 + · · ·+ δlωl) ∈ δKΩ∗(P )[δ]

d∗δ(ω + δω1 + · · ·+ δlωl) ∈ δKΩ∗(P )[δ] .

This forms a spectral sequence which is isomorphic to the Leray–Serre spectral

sequence for the fibration G → Pπ→ M . In addition to the abstract convergence

Ek∞∼= Hk(P ; R) (where k = i+ j) given by the Leray–Serre spectral sequence, there

is the geometric/analytic statement that

Lk0

def= lim

δ→0KerLk

gδ= Ek

∞.

In fact, the space Lk0 is a smooth extension of KerLk

gδto δ = 0. This also gives a

smooth extension of Ker ∆kgδ

to δ = 0. If we define the formal rescaled Laplacian

Lkδ : Ωp(P )[|δ|] −→ Ωp(P )[|δ|],

then elements ωδ ∈ KerLδ are given by the Taylor series at δ = 0 of sections

C∞([0, 1],KerLkgδ

), where KerLkgδ

is a finite-dimensional vector bundle over [0, 1].

The fact that Ei,jK is isomorphic to the Leray–Serre spectral sequence allows us

to write down the lower-order terms (with respect to δ) in a basis of KerLδ. A priori,

we would have to construct a power series that is formally harmonic. However, if

11

Page 19: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

we have a polynomial ωδ = ω +O(δ) ∈ Ωk(P )[δ] such that

dδωδ ∈ δN(k)Ωk+1(P )[δ], d∗δωδ ∈ δN(k)Ωk−1(P )[δ],

where N(k) is the term where the spectral sequence calculating Hk(P ; R) collapses,

then we are guaranteed the existence of higher order terms making ω+O(δ) formally

Lδ-harmonic. In fact, the same argument implies that if we have

ωδ = ω + δω1 + · · ·+ δkωk +O(δk+1) ∈ Ωk(P )[δ]

such that

dδωδ ∈ δN(k)+kΩk+1(P )[δ], d∗δωδ ∈ δN(k)+kΩk−1(P )[δ],

then there exists a power series of the form

ω + δω1 + · · ·+ δkωk +O(δk+1) ∈ KerLδ.

Furthermore, if this power series is in the domain of ρ−1δ (acting on formal power

series), then we see the following relationship with ∆gδ-harmonic forms in the adia-

batic limit:

ω0,k + ω1,k−11 + · · ·+ ωk,0

k ∈ Ker ∆k0 = lim

δ→0Ker ∆k

where ωi,jl denotes the projection of ωl onto Ωi,j(P ). This is the method by which

we describe Ker ∆k0 in calculations.

Chapter 5 is the technical heart of the thesis. Roughly speaking, we look at

principal G-bundles where the first real characteristic class is 0 and prove that

low-dimensional harmonic forms (k = 1, 2, 3) on P in the adiabatic limit have a

natural description in terms of a Chern–Simons form on P and harmonic forms on

(M, g). This is usually done in two parts. First, we describe the lower-order terms

of elements in KerLδ. The vanishing of the relevant characteristic class implies

that the relevant part of the spectral sequence collapses at N = 2. Therefore, we

12

Page 20: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

only need to produce polynomials in Ω∗(P ) such that dδ and d∗δ are of the order

of δ2+k (where k = 1, 2, 3). Our explicit description of the right-invariant forms on

P , along with standard properties of the Chern–Simons forms, makes it possible

to produce such polynomials. After describing KerLδ, we use the isometry ρ−1δ to

obtain elements of Ker ∆0. We frequently use the notation

Hkg(M)

def= Ker ∆k

g ⊂ Ωk(M)

to denote the harmonic k-forms on the base with respect to the metric g.

The first case is for a principal U(n)-bundle P with connection Θ over a Rieman-

nian manifold (M, g). We denote this by (M,P, g,Θ). If c1(P ) = 0 ∈ H2(M ; R),

then for any δ ≥ 0 (where gG is any Ad-invariant metric on u(n)),

Ker ∆1gδ

= R[i

2πTr(Θ)− π∗h]⊕ π∗H1

g(M).

Here, h ∈ Ω1(M) is the unique form (under the Hodge decomposition) such that

h ∈ d∗Ω2(M) and dh = c1(P,Θ). The second case is for G semi-simple. Given

(M,P, g,Θ) and any Ad-invariant metric gG on g,

Ker ∆1gδ

= π∗H1g(M) (∀δ ≥ 0),

Ker ∆20 = π∗H2

g(M).

These two cases are included primarily as a warm-up to the next case.

The main technical result of the thesis is Theorem 5.3.2. For G simple, we begin

with a principal G-bundle P with connection Θ over a Riemannian manifold (M, g).

Letting Θ denote the curvature 2-form, we assume that the real characteristic class

[〈Ω ∧ Ω〉] = 0 ∈ H4(M ; R)

13

Page 21: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

(e.g. p1

2(P ) = 0 or c2(P ) = 0 in H4(M ; R)). Then, for any Ad-invariant gG,

Ker ∆30 = lim

δ→0Ker ∆gδ

= R[α(Θ)− π∗h]⊕ π∗H3g(M).

Here, α(Θ) is the Chern–Simons 3-form, and h ∈ Ω3(M) is the unique form such

that h ∈ d∗Ω4(M) and dh = 〈Ω ∧ Ω〉.

In Chapter 6, we give a topological definition of a string structure. Note that

(for n ≥ 5)

π3(Spin(n)) ∼= H3(Spin(n); Z) ∼= Z.

Up to homotopy, we define the topological group String(n) as the 3-connected cover

of Spin(n). The existence of such a group is given by [ST], but we do not need any

other properties of these more explicit descriptions. For a Spin(n)-bundle Pπ→M ,

there is a 1-1 correspondence between isomorphism classes of String(n)-bundles

covering P and cohomology classes S ∈ H3(P ; Z) such that i∗S ∈ H3(Spin(n); Z) is

the (preferred) generator. We define string structures as these cohomology classes. A

string structure exists if and only if p1

2(P ) = 0 ∈ H4(M ; Z), and the space of string

structures is a torsor for H3(M ; Z). This is actually discussed for more general

situations, and then specialized to define spin, U , spinc, and string structures.

This leads to the following construction: given a principal Spin(n)-bundle P

with connection Θ over (M, g), along with a string structure S ∈ H3(P ; Z), take the

harmonic representative in the adiabatic limit of the image of S in real cohomology.

As a consequence of Theorem 5.3.2, this is of the form (Theorem 6.4.4)

[S]0 = limδ→0

[S]gδ= α(Θ)− π∗Hg,Θ,S ∈ Ω3(P ).

The choice of metric, connection, and string structure give a canonical 3-formHg,Θ,S.

14

Page 22: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Letting AP denote the affine space of connections on P , the map

Met(M)×AP × String Structures −→ Ω3(M)

is equivariant with respect to the action of H3(M ; Z), which acts on the space of

string structures by adding the pulled back cohomology class, and which acts on

Ω3(M) by adding the harmonic representative of a class. In the above construction,

dHg,Θ,S = p1

2(P,Θ). Furthermore, Hg,Θ,S and the pulled back Chern–Simons form

give the same values when evaluated on any 3-cycle in M . (The choice of string

structure gives a global section, up to homotopy, on any 3-cycle in M .)

In particular, if Spin(M) → M is the frame bundle for a Riemannian manifold

with spin structure, the metric induces the Levi-Civita connection. Therefore, we

have a map

Met(M)× String Structures −→ Ω3(M)

which is equivariant under the natural action of H3(M ; Z). Analogous statements

also hold for U and spinc structures (Theorems 6.2.5 and 6.3.3).

In Chapter 7, we use this construction to define canonical metric connections

associated to string structures. Given (M, g, S), we obtain the canonical 3-form

Hg,S and define a metric connection ∇g,S on TM whose torsion is given by

〈T g,S(X, Y ), Z〉g = Hg,S(X, Y, Z).

Though the form Hg,S is independent of a global rescaling on M , the resulting

connection uses the metric g to define the torsion tensor, and hence it is natural

to introduce the parameter ε > 0. This gives a 1-parameter family of connections

∇εg,S, which preserve the metric g and have torsion 1εT g,S. If ∇g is the Levi-Civita

15

Page 23: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

connection, then the Ricci tensor for ∇εg,S is given by

Ricεg,S(X, Y ) = Ricg(X,Y ) +1

∑i

〈(∇geiT )(X, Y ), ei〉 −

1

4ε2

∑i

〈TeiX,Tei

Y 〉.

16

Page 24: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 2

PRELIMINARIES

The main theorems in this thesis involve principal bundles P , Hodge theory,

and Chern–Simons forms. In this chapter, we introduce these concepts and the

properties which we use. In particular, since we later restrict to the subcomplex

of G-invariant forms on a principal bundle, we set up the description of TP as

a G-equivariant bundle. Along the way we introduce the notion of Lie algebra

cohomology and are always careful to avoid confusions between left and right.

2.1 Lie groups and torsors

For simplicity, we always deal with G a compact Lie group, and use Lg to denote

the isomorphism given by left multiplication of g ∈ G and Rg for right multiplication

by g ∈ G.

Lg : G→ G Rg : G→ G

g1 7→ g · g1 g1 7→ g1 · g

These induce maps on the tangent spaces

(Lg)∗ : Tg1G→ Tg·g1G, (Rg)∗ : Tg1G→ Tg1·gG,

and

(Adg)∗def= (Lg)∗ (Rg−1)∗ : Tg1G→ Tg·g1·g−1G.

17

Page 25: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

We denote the induced map on the tangent bundles by

(Lg)∗ : TG→ TG, (Rg)∗ : TG→ TG,

and the induced map on smooth vector fields

(Lg)∗ : C∞(G, TG)→ C∞(G, TG), (Rg)∗ : C∞(G, TG)→ C∞(G, TG).

Definition 2.1.1. To any Lie group G we canonically associate the Lie algebra g

of left-invariant vector fields:

gdef= left-invariant vector fields on G = V ∈ C∞(G, TG)|(Lg)∗V = V ∀g ∈ G.

In other words, an element V ∈ g is a map V : G → TG such that V (g1) ∈ Tg1G

for every g1, and

(Lg)∗V (g1) = V (g · g1)

for every g ∈ G. The bracket [·, ·] is defined by the usual Lie bracket on vector fields.

Definition 2.1.2. The Adjoint action of G on g is defined by

Adg : g −→ g

V 7→ (Lg)∗(Rg−1)∗V = (Rg−1)∗V.

Given any vector in the tangent space of a point, we can associate to it a left-

invariant vector field by left-translating the vector; i.e.

(LG)∗ : Tg1G→ g ⊂ C∞(G, TG)

v 7→ ((Lgg1 7→ (Lg)∗v) .

18

Page 26: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Definition 2.1.3. The Maurer-Cartan 1-form θ ∈ Ω1(G; g) (often denoted g−1dg),

is the g-valued 1-form on G defined by associating to a tangent vector the vector

field obtained by left-multiplication; i.e.

θ : TG→ g

(g, v) 7→ (LG)∗v.

Proposition 2.1.4. The Maurer-Cartan 1-form θ ∈ Ω1(G; g) is left-invariant, and

is right-equivariant:

L∗gθ = θ, R∗gθ = Adg−1θ.

It also satisfies the equation

dθ = −1

2[θ ∧ θ].

Proof. For any vector field V ∈ g, (Lg)∗V = V and (Rg)∗V = Adg−1V . For an

arbitrary vector v ∈ Tg′G, by definition θ(v) = (LG)∗v ∈ g. Therefore,

L∗gθ(v) = θ((Lg)∗v) = (LG)∗(Lg)∗v = (LG)∗v = θ(v),

and

R∗gθ(v) = θ((Rg)∗v) = (LG)∗(Rg)∗v = Rg∗(LG)∗v = Adg−1θ(v).

To evaluate the derivative, it suffices to use left-invariant vector fields V0, V1 on G.

Then, as a function with values in a fixed vector space g, θ(Vi) is constant. Hence,

dθ(V0, V1) = V0θ(V1)− V1θ(V0)− θ([V0, V1])

= −θ([V0, V1]) = −1

2([θ(V0), θ(V1)]− [θ(V1), θ(V0)])

= −1

2[θ ∧ θ](V0, V1).

19

Page 27: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Now, we wish to discuss right G-torsors. A torsor for a group G is a space Y on

which G acts freely and transitively. Torsors appear frequently, and in our case the

fibers of a principal bundle are torsors.

Suppose Y is a smooth manifold that has a smooth right action of G which is

free and transitive

Y ×G→ Y.

Choosing a point y0 ∈ Y gives us the diffeomorphism ϕy0 : Y → G by

Yϕy0−→ G

y = y0 · g 7−→ g.

Note that ϕy0 is equivariant with respect to right multiplication by G. Choosing a

different point y1 ∈ Y gives another diffeomorphism which differs by left multipli-

cation in G. In other words, since y0 = y1 · gy1y0 , the following diagram commutes:

g_

Lgy1y0

y0 · g = y1 · gy1y0 · g+

ϕy0

55kkkkkkkkkkkkkkkk

ϕy1

))SSSSSSSSSSSSSS

gy1y0 · g.

(2.1.1)

Therefore, ϕy1 ϕ−1y0

= Lgy1y0: G → G. This implies that any structure on G

which is invariant under left multiplication induces a structure on the torsor Y . In

particular, the Maurer-Cartan 1-form on θ ∈ Ω1(G; g) pulls back to a Lie-algebra

valued 1-form on the torsor Y . We will also call this form θ ∈ Ω1(Y ; g).

Proposition 2.1.5. If Y is a right G-torsor, there is a canonical θ = ϕ∗yθ ∈

Ω1(Y ; g). This induces the canonical G-equivariant isomorphism

TY ∼= Y × g.

20

Page 28: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proof. Choosing some point y0 ∈ Y gives ϕ∗y0θ ∈ Ω1(Y ; g). A choice of a different

point y1 differs only by left multiplication in G. By the left-invariance of θ (Propo-

sition 2.1.4) and the fact that ϕy0 and ϕy1 differ by left-multiplication (2.1.1),

ϕ∗y1θ = ϕ∗y0

L∗gy1y0θ = ϕ∗y0

θ.

Therefore, we have a well-defined g-valued 1-form on Y which we will also denote

θ. At each tangent space, we then see the canonical vector space isomorphism

θ : TyY≈−→ g

v 7−→ θ(v)

and therefore the isomorphism

TYY×θ−→ Y × g.

By Proposition 2.1.4, θ((Rg)∗v) = Adg−1θ(v). Hence, this isomorphism is equivari-

ant with respect to the right G action on TY and g.

We now define the abstract vector space

gYdef= Y ×Ad g = (Y × g) / ((y, v) ∼ (yg, Adg−1v)) . (2.1.2)

It is non-canonically isomorphic to g. Once we choose a point y ∈ Y , then we have

an isomorphism given by the fact that any element of gY can be written as (y, v)

for v ∈ g. If we consider this vector space as a vector bundle over a point

gY −→ pt,

and we have the map Yπ→ pt, then we have the following equivalence of bundles.

Proposition 2.1.6. There are canonical isomorphisms of G-equivariant vector bun-

21

Page 29: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

dles

TY ∼= Y × g ∼= π∗gY .

Proof. The first isomorphism was given in Proposition 2.1.5. To see the second,

note that by definition,

π∗gY |y = (y, e) | e ∈ gY .

Using the chosen y, we have the canonical isomorphism gY∼= g. Therefore,

π∗gY∼= Y × g.

2.2 Right invariant vector fields and forms on a G-torsor

What do the right-invariant vector fields on a torsor, denoted C∞(Y, TY )G, look

like? Using the isomorphism in Proposition 2.1.6, we have the following equivalence:

C∞(Y, TY )G ∼= C∞(Y, Y × g)G ∼= C∞(Y, π∗gY )G ∼= π∗C∞(pt, gY ).

More concretely, we have natural equivalences between right-invariant vector fields,

elements of gY , and functions of the form

V : Y → g such that V (yg) = Adg−1V (y).

When using this translation, for a given vector V ∈ gY we will denote the cor-

responding vector field by V : Y → g. Because the Lie bracket [·, ·] on g is Ad

equivariant (or GAd→ Gl(g) is a representation), the induced bracket

(Y × g)× (Y × g)[·,·]−→ (Y × g)

(p, V1)× (p, V2) 7−→ (p, [V1, V2])

22

Page 30: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

is Ad-equivariant and descends to

gY × gY[·,·]−→ gY . (2.2.1)

More concretely, the bracket is defined by

˜[V1, V2](y) = [V1(y), V2(y)]. (2.2.2)

Lemma 2.2.1. The Lie bracket of right-invariant vector fields on Y is given by the

bracket on gY from (2.2.1); i.e. the following diagram is commutative:

gY × gY

[·,·] // gY

C∞(Y, TY )G × C∞(Y, TY )G[·,·] // C∞(Y, TY )G.

Proof. Let V1, V2 ∈ gY , and V1, V2 : Y → g the induced vector fields in C∞(Y, TY )G.

Because the Vi are right-invariant, around any point y ∈ Y we have that Vi(yg) =

Adg−1Vi(y). Therefore, the bracket [V1, V2](y) ∈ g is given by

[V1, V2](y) = [Adg−1V1(y), Adg−1V2(y)](y) =(Adg−1 [V1(y), V2(y)]

)(y) = [V1(y), V2(y)].

The middle equality follows from the Ad-equivariance of the Lie bracket. By defi-

nition of the bracket on gY (2.2.2), the above diagram is commutative.

To discuss right-invariant differential forms, we define the vector spaces

Λkg∗Ydef= Y ×Ad Λkg∗ = Y × Λkg∗/

((y, ψ) ∼ (yg, Ad∗g−1ψ)

).

Then, we have canonical isomorphisms

Ωk(Y )G ∼= C∞(Y, Y × Λkg)G = π∗C∞(pt,Λkg∗Y ). (2.2.3)

23

Page 31: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This gives us an isomorphism of cochain complexes

Ωk(Y )G, d ∼= Λkg∗Y , dg,

where dg is at this point only defined by d and the isomorphism (2.2.3). We now

wish to give a more intrinsic definition of dg. Let ψ ∈ Λkg∗Y , V0, . . . , Vk ∈ gY , and

ψ = π∗ψ, Vi = π∗Vi. Then,

dgψ(V0, . . . , Vk) =dψ(V0, . . . , Vk)

=∑

i

(−1)iViψ(V0, . . . ,Vi, . . . , Vk)

+∑i<j

(−1)i+jψ([Vi, Vj], . . . ,Vi, . . . ,

Vj, . . . , Vk).

Because ψ and Vi are right-invariant, then ψ(V0, . . . ,Vi, . . . , Vk) is a constant function

on Y , and its derivative is 0 along any vector field. Therefore,

dgψ(V0, . . . , Vk) =∑i<j

(−1)i+jψ([Vi, Vj], . . . , Vi, . . . , Vj, . . . , Vk).

This gives an intrinsic definition of dg in terms of the Lie bracket that coincides

with the exterior derivative of right-invariant forms; i.e. the following diagram is

commutative:

Λkg∗Y

π∗

dg // Λk+1g∗Y

π∗

Ωk(Y )G d // Ωk+1(Y )G.

(2.2.4)

A classical theorem of Chevalley and Eilenberg shows that the inclusion of

cochain complexes

right-invariant forms on Y → Ω∗(Y )

induces an isomorphism on the cohomology of these complexes (see Theorem 2.3 in

24

Page 32: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

[CE]). Therefore, to compute the de Rham cohomology of Y , it suffices to look at

the finite dimensional complex of right-invariant forms, which is much easier than

looking at the infinite dimensional complex of arbitary forms. As seen above, it

is easy to explicitly describe the derivative dg via information encoded in the Lie

algebra g. In fact, the first and last dg are both trivial.

Lemma 2.2.2. For Y a right G torsor, the derivatives dg : R → g∗Y and dg :

Λn−1g∗Y → Λng∗Y are both 0 (where n = dim g).

Proof. Note that H0(g) and Hn(g) are both isomorphic to R (G is compact) and

dim Λ0g∗Y = dim Λng∗Y = 1.

Therefore, Ker (dg : Λ0g∗Y → Λ1g∗Y ) = Λ0g∗Y , and Image(dg : Λn−1g∗Y → Λng∗Y ) = 0.

Therefore, the cochain complex of right-invariant forms is canonically isomorphic

to

0→ R 0→ g∗Ydg→ Λ2g∗Y

dg→ · · · dg→ Λn−1g∗Y0→ Λng∗Y → 0.

2.3 Connections on principal G-bundles

Again, we assume G to be a compact Lie group. Informally, a principal G-bundle

P is a family of right G torsors parameterized by a space M . G acts (from the right)

freely and transitively on each fiber. The previous constructions on right G-torsors

were canonical and will induce the same constructions on a principal bundle. See

section 1 of [Fre1] for another brief overview of connections on principal bundles.

Details and explicit proofs are found in sections II.2, II.3, II.5, and III.1 of [KN].

Definition 2.3.1. A principal G-bundle Pπ→ M is a manifold P with a free right

G action. Then M = P/G is also a manifold.

25

Page 33: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Example 2.3.2. The usual example of a principal G-bundle is the frame bundle for

some vector bundle. Assume that G ⊂Mn(F) (where F = R or C) is defined as the

group of linear transformations which preserve some structure on Fn. For example,

O(n) is the group of linear transformations which preserves the standard metric on

Rn, SO(n) preserves the metric and orientation, and U(n) preserves the standard

hermitian metric on Cn. Then, given a vector bundle E → M equipped with said

structure, define the set of structure preserving frames on Ex

G(Ex) = Fn f→ Ex|f is compatible with structure.

G(Ex) is a right torsor for G, where the action is given by precomposing with g :

Fn → Fn. The spaces G(Ex) are parameterized by M , and taken together form a

principal bundle G(E)π→M .

Let Pxdef= π

−1(x) denote the fiber at a point x ∈ M . Because G acts freely on

P , then Px has a free and transitive right G action. In other words, Px is a right

G-torsor. Define the distribution of vertical tangent vectors by

T V Pdef= Ker(π∗) ⊂ TP.

At any point p ∈ P , where π(p) = x,

T V P|p = Ker(π∗)|p = T (Ker π)|p = (TPx)|p.

Since Px is a G-torsor, then the Maurer-Cartan 1-form θ ∈ Ω1(Px, g) gives the

canonical G-equivariant isomorphism (Proposition 2.1.5)

TPx∼= Px × g

(p, v) 7→ (p, θ(v)).

26

Page 34: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

We define the adjoint bundle gP →M as

(gP →M)def= P ×Ad g = (P × g) / ((pg, Adg−1v) (p, v)) .

This is simply a parameterized version of the vector space in equation (2.1.2). Since

the isomorphisms in Proposition 2.1.6 were canonical, we then have the isomor-

phisms of G-equivariant vector bundles

T V P ∼= P × g ∼= π∗gP . (2.3.1)

Though there is a canonical vertical subspace at each point, one must make a choice

in picking out a horizontal subspace. This is due to the that the fibers of P are not

canonically isomorphic to G, and so there is no intrinsic way to identify separate

fibers with each other. This problem is resolved by introducing a connection on

P , which is really just an equivariant projection from TP onto the vertical tangent

bundle T V P .

Definition 2.3.3. A connection on P →M is a g-valued 1-form Θ ∈ Ω1(P ; g) suchthat

1. i∗xΘ = θ ∈ Ω1(Px; g),

2. R∗gΘ = Adg−1Θ,

where ix : Px → P is the fiber-wise inclusion.

The first condition is equivalent to Θ : TpP → g ∼= T Vp P being the identity

map on vertical vectors. The second condition implies that the projections are

equivariant, since Θ(Rg∗v) = Adg−1Θ(v) for any vector.

Although the vertical subspaces in P are canonically defined, it is the connection

Θ ∈ Ω1(P ; g) that picks out a horizontal subspace at each point p ∈ P . This gives

a right-equivariant distribution

THPdef= Ker(Θ) ⊂ TP, (2.3.2)

27

Page 35: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

and THP ⊕ T V P = TP. A connection Θ then gives the canonical isomorphism

π∗ : THp P

≈−→ Tπ(x)M. (2.3.3)

This induces the canonical isomorphism at every point p ∈ P ,

π∗TMp = v ∈ TMπ(p)π∗∼= v ∈ THPp.

and hence the canonical isomorphism of G-equivariant vector bundles

THP ∼= π∗TM. (2.3.4)

A connection also gives a way to identify separate fibers by parallel translation.

First choose a path

γ : [0, 1]→M,

but we only consider [0, 1] as a smooth manifold. We have not defined a metric

on [0, 1] (but use the numbers for notational convenience). Then dγ : T [0, 1] →

TM gives a 1-dimensional distribution in TM, which we refer to as dγ. Using

the isomorphism induced by the connection in (2.3.3), the distribution dγ lifts to

a 1-dimensional distribution dγ ⊂ TP|π−1(γ). Any 1-dimensional distribution is

integrable, so for any point p0 ∈ Pγ(0) there exists a unique integral submanifold

γp0 : [0, 1]→ P that is everywhere tangent to the distribution and γp0(0) = p0. The

horizontal distribution THP is equivariant, and hence we have a G-equivariant map

called parallel translation

‖γ: Pγ(0) −→ Pγ(1)

p0 7−→ γp0(1).

This construction does not involve a metric on [0, 1] and therefore is invariant under

diffeomorphisms of [0, 1], or reparameterizations of the path γ.

28

Page 36: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

The horizontal subspace also induces covariant differentiation in associated vec-

tor bundles. Let W be a G representation, denoted ρ : G→ Gl(W ). The associated

vector bundle is defined by

WP = P ×G W = P ×W/((p, w) (pg, ρ(g−1)w)

)and sections C∞(M,WP ) are equivalent to equivariant functions P → W . In order

to take the derivative of a function, its values must live in a fixed vector space, and

sections of a vector bundle live in a parameterized family of vector spaces. However,

the values of functions P → W live in a fixed vector space. Let ξ ∈ C∞(M,WP )

be a section of the vector bundle and ξ ∈ C∞(P,W ) the associated equivariant

function. Given a vector field X ∈ C∞(M,TM), there is a unique horizontal vector

field X ∈ C∞(P, THP ) such that π∗X = X. It then makes sense to derivate ξ along

the vector field X

d eX ξ ∈ C∞(P,W ).

Because the function f and the vector field X are G-equivariant, d eX ξ is also G-

equivariant. Define ∇Xξ ∈ C∞(M,PW ) to be unique section such that

∇Xξ = d eX ξ. (2.3.5)

The operator ∇ : C∞(M,EP )→ C∞(M,T ∗M ⊗EP ) is called the connection on the

associated vector bundle. We also denote this

Ω0(M ;WP )∇−→ Ω1(M ;WP ).

It is important to remember that ∇ depends on the choice of a connection on P in

order to lift the vector field on M to a vector field on P . For f ∈ C∞(M), ∇ also

satisfies the Leibniz rule

∇(fξ) = (df)ξ + f∇ξ.

29

Page 37: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This is seen by lifting fξ to f ξ and derivating. We can then extend ∇ as first order

differential operator to the complex

Ω0(M ;WP )∇−→ Ω1(M ;WP )

d∇−→ Ω2(M ;WP )d∇−→ Ω3(M ;WP )

d∇−→ · · · (2.3.6)

by requiring that

d∇(ω ∧ ξ) = (dω) ∧ ξ + (−1)iω ∧ (∇ξ)

for ω ∈ Ωi(M) and ξ ∈ C∞(M,WP ).

The possibility that THP is not an integrable distribution leads to the notion of

curvature.

Definition 2.3.4. Given a connection Θ ∈ Ω1(P ; g), the curvature 2-form Ω ∈

Ω2(P ; g) is

Ωdef= dΘ +

1

2[Θ ∧Θ].

Proposition 2.3.5. R∗gΩ = Adg−1Ω, and Ω is only non-zero when evaluated on

horizontal vectors.

Proof. G equivariance of Ω follows from the equivariance of Θ.

The Maurer-Cartan equation in Proposition 2.1.4 and the fact that i∗xΘ = θ for

the inclusion ix : Px → P imply

i∗xΩ = dθ +1

2[θ ∧ θ] = 0.

Therefore, Ω is zero when evaluated on two vertical vectors.

Let XH be a horizontal vector and XV a vertical vector. Then

Ω(XH , XV ) = XHΘ(XV )−XV Θ(XH)−Θ([XH , XV ]).

By definition, Θ(XH) = 0 and Θ(XV ) = XV . Viewing XV as a function P → g, we

30

Page 38: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

see that

XHΘ(XV )−Θ([XH , XV ]) = dXHXV − dXH

XV = 0.

Since Ω is a G-equivariant, purely horizontal 2-form on P , we can view it as an

element of Ω2(M ; gP ). From the definition, we see that given two horizontal vector

fields VH0 , VH1 on P ,

Ω(VH0 , VH1) = dΘ(VH0 , VH1) = −Θ([VH0 , VH1 ]). (2.3.7)

In other words, the curvature is (up to sign) given by taking the Lie bracket of two

horizontal vector fields and then projecting onto the fiber. The curvature measures

the non-integrability of the distribution THP . A connection is called flat if Ω = 0,

or equivalently, if the distribution THP is integrable.

2.4 Hodge theory

de Rham cohomology allows us to view real cohomology in a more geometric way

by interpreting cohomology classes in terms of differential forms. However, a given

cohomology class only determines a closed differential form up to an additional exact

form. The basic idea of Hodge theory is that the introduction of a metric canonically

picks out a particular differential form by requiring that the form also be coclosed

with respect to an adjoint differential.

Let (V, g) be an n-dimensional, oriented, Euclidean vector space. The Euclidean

metric g induces a metric 〈·, ·〉 on ⊕kΛkV . This also gives an isomorphism ΛnV ∼= R,

and we denote ω the element which maps to 1. (When V = TxM∗, then ω = volx.)

This gives rise to an isomorphism, called the Hodge star

∗ : ΛkV → Λn−kV,

31

Page 39: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

where, for γ ∈ ΛkV , ∗γ is the unique element determined by

γ ∧ β = 〈∗γ, β〉ω

for all β ∈ Λn−kV . Useful properties include

• ∗(∗γ) = (−1)k(n−k)γ,

• ∗1 = ω.

Let (M, g) be a compact oriented n-dimensional Riemannian manifold without

boundary. By the above construction, for every x ∈M we obtain

∗ : ΛkTxM∗ → Λn−kTxM

∗,

and therefore we have an isomorphism of smooth differential forms

∗ : Ωk(M)→ Ωn−k(M).

Define d∗ : Ωk(M)→ Ωk−1(M) by

d∗def= (−1)n(k+1)+1 ∗ d ∗ .

An integration by parts argument shows that d∗ is the adjoint for d. Namely, if (·, ·)

is the inner product on differential forms induced by integrating the pointwise inner

product, then

(dγ, β) = (γ, d∗β). (2.4.1)

In other words, the metric picks out an adjoint d∗ to d and gives

0d // Ω0(M)

d∗

d // Ω1(M)d //

d∗uu

Ω2(M)d //

d∗uu

· · · d //

d∗vv

Ωn−1(M)d //

d∗

xxΩn(M) d //

d∗tt

0.

d∗vv

The Hodge star gives an isomorphism between the cochain complex (Ω∗(M), d) and

32

Page 40: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

its dual complex (Ω∗(M), d∗), and therefore

Hk(Ω∗(M), d∗) ∼= Hn−k(Ω∗(M), d). (2.4.2)

Now, define the Hodge Laplacian

∆k def= (d+ d∗)2 = dd∗ + d∗d : Ωk(M)→ Ωk(M).

In fact, d + d∗ is formally self-adjoint, and therefore Ker(d + d∗)2 = Ker(d + d∗).

Since d increases the degree of a form while d∗ decreases the degree,

Ker ∆k = Ker d ∩Ker d∗ ⊂ Ωk(M).

We denote Hkg(M) = Ker ∆k ⊂ Ωk(M) to be the harmonic k-forms.

Theorem 2.4.1. (Hodge) Let (M, g) be a compact, oriented, Riemannian manifold

without boundary. Then there is an orthogonal decomposition

Ωk(M) = dΩk−1(M)⊕ d∗Ωk+1(M)⊕Hkg(M).

Consequently, in each de Rham class [β] there is a unique harmonic representative,

and Hkg(M) ∼= Hk(M ; R).

Corollary 2.4.2. Any form ω ∈ d∗Ωk+1(M) is uniquely determined by dω. Any

form ω ∈ dΩk−1(M) is uniquely determined by d∗ω.

Proof. Since d∗ is the adjoint to d, the corollary follows from the isomorphisms

d∗Ωk+1

d11 dΩ

k(M).d∗qq

We treat one case more concretely. Suppose ω1, ω2 ∈ d∗Ωk+1(M), and dω1 = dω2.

Then, ω1−ω2 is closed, and hence in dΩk−1(M)⊕Hkg(M). But, ω1−ω2 ∈ d∗Ωk+1(M),

and hence ω1 = ω2.

33

Page 41: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

It should be noted that while d and d∗ are both locally defined, solving ∆ψ = 0

is a global problem and in general very difficult. However, if a group G acts by

isometries on M , then this problem is more tractable. First, note that because ∗

is defined solely in terms of the oriented metric, then ∗ commutes with isometries.

Therefore, the Hodge star maps the subspace of G-invariant forms to itself, and ∆

commutes with isometries. By the following proposition, when finding harmonic

forms it suffices to look at the subcomplex of G-invariant forms.

Proposition 2.4.3. Suppose that G is a connected Lie group which acts on (M, g)

by isometries. If ψ ∈ Hkg(M), then ψ is invariant under the G action; i.e. ψ ∈

Ωk(M)G.

Proof. Let γ : [0, 1]→ G be a path connecting the identity e to any element h ∈ G.

Then, since e∗ψ = ψ, and ∆ ((γ(t))∗ψ) = γ(t)∗∆ψ = 0, we have a path in the space

Hkg(M). The forms γ(t)∗ψ all represent the same cohomology class, so this path

must be constant.

In particular, if G is a compact Lie group equipped with a bi-invariant metric,

then any harmonic form must be bi-invariant. At the Lie algebra level, this implies

that any harmonic form ψ ∈ Λkg must be Ad-invariant. Conversely, if ψ is an

Ad-invariant form, then it must be closed, since

0 =d

dt

∣∣∣∣t=0

(Ad∗etX0ψ(X1, . . . , Xk)

)=

d

dt

∣∣∣∣t=0

ψ(AdetX0X1, . . . , AdetX0Xk)

=d

dt

∣∣∣∣t=0

ψ(AdetX0X1, X2, . . . , Xk) + · · ·+ d

dt

∣∣∣∣t=0

ψ(X1, X2, . . . , AdetX0Xk)

= ψ([X0, X1], X2, . . . , Xk) + · · ·+ ψ(X0, . . . , [X0, Xk])

= dψ(X0, X1, . . . , Xk).

(The third equality follows from the linearity of ψ.) Furthermore, if the metric on g

34

Page 42: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

is Ad-invariant, then the Hodge star of Ad-invariant forms will be Ad-invariant and

hence closed.

Proposition 2.4.4. For an Ad-invariant metric on g, The harmonic forms on the

finite-dimensional complex

0 // R 0 //~~

g∗dg //

0

Λ2g∗dg //

d∗gzz

· · · dg //

d∗gww

Λn−1g∗0 //

d∗g

yyΛng∗ //

0vv

0yy

are precisely the Ad-invariant forms. This also implies that for G a compact and

connected Lie group with a bi-invariant metric, the harmonic forms are precisely the

bi-invariant forms.

We later deal with rescalings on the metric, so we prove the following useful

lemma.

Lemma 2.4.5. If gδ = δ2g, then ∗gδ= δ2k−n∗g : ΛkV → Λn−kV .

Proof. The metric 〈·, ·〉 on Λn−kV is induced by

〈v1 ∧ . . . ∧ vn−k, w1 ∧ . . . ∧ wn−k〉g = Det (g(va, wb)) ,

where (g(va, wb)) denotes an (n− k)× (n− k) matrix. Then 〈·, ·〉δ2g = δ2(n−k)〈·, ·〉g.

The isometry (V, g)→ (V, gδ) sends v 7→ δ−1v, and therefore

ωgδ= δ−nωg.

Thus,

γ ∧ β = 〈∗gδγ, β〉gδ

ωgδ= δn−2k〈∗gδ

γ, β〉gωg

= 〈∗gγ, β〉gωg,

and so ∗gδ= δ2k−n∗g.

35

Page 43: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

2.5 Chern–Weil and Chern–Simons Forms

In Chapter 5, we only look at the Chern–Simons 1- and 3-forms, so the following

treatment will be very concrete. The theory of Chern–Weil and Chern–Simons

forms is quite general, though, and most of the properties stated here generalize.

The paper [CS] contains these details, and sections 1.2-1.4 of [Fre2] contain an

excellent summary of Chern–Weil and Chern–Simons forms. Section 1 of [Fre1] also

proves the important properties of the Chern–Simons 3-form. In general, a Chern–

Weil form lives in Ω2k(M), and when pulled back to P , it is the derivative of a

Chern–Simons form in Ω2k−1(P ).

Definition 2.5.1. Given a U(n)-bundle Pπ→M with connection Θ, the first Chern-

form is

c1(P,Θ)def=

i

2πTr(Ω) ∈ Ω2(M).

where Tr is the trace of the standard matrix representation of the Lie algebra u(n).

The form c1(P,Θ) is a Chern–Weil form. Of course, Ω ∈ Ω2(M ; gP ) takes values

in the Adjoint bundle. However, since Tr is an Ad-invariant linear map g→ R, we

get a well-defined 2-form on M with values in R. In general, Chern–Weil forms are

obtained from applying an Ad-invariant polynomial to the curvature 2-form.

Definition 2.5.2. The Chern–Simons 1-form of a U(n)-bundle with connection

(P,Θ) over M is

i

2πTr(Θ) ∈ Ω1(P ).

Proposition 2.5.3. i2πTr(Θ) ∈ Ω1(P ) satisfies the following properties.

1. i∗x(i

2πTr(Θ)) = i

2πTr(θ) ∈ Ω1(G) is bi-invariant.

2. [ i2πTr(θ)] ∈ H1(U(n); R) is the R-image of the standard generator in H1(U(n); Z) ∼=

Z.

36

Page 44: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

3. R∗g(

i2πTr(Θ)) = i

2πTr(Θ); i.e. i

2πTr(Θ) is right-invariant.

4. d i2πTr(Θ) = i

2πTr(Ω) = π∗c1(P,Θ).

Proof. By definition, i∗xΘ = θ, so i∗x(i

2πTr(Θ)) = i

2πTr(θ). The Ad-invariance of

trace implies the bi-invariance of i2πTr(θ). In particular, this means that i

2πTr(θ)

is closed and represents a de Rham class.

To see the second property, first note that the inclusion U(1) → U(n) induces

an isomorphism on the first cohomology. [ i2πTr(θ)] is the real class representing the

R-image of the standard generator for H1(U(1); Z), as seen by∫U(1)

i

2πTr(θ) = 1.

The third property follows from the right-equivariance of Θ and theAd-invariance

of trace, since

R∗g(i

2πTr(Θ)) =

i

2πTr(Adg−1Θ) =

i

2πTr(Θ).

Finally,

di

2πTr(Θ) =

i

2πTr(dΘ) =

i

2πTr(dΘ +

1

2[Θ ∧Θ]) =

i

2πTr(Ω) = π∗c1(P,Θ).

The middle equality follows from Tr([Θ ∧Θ]) = 0.

We now deal with the situation where G is a compact simple group (the main

groups of interest will be Spin(n)).

Definition 2.5.4. Given a principal G-bundle Pπ→ M with connection Θ, where

G is simple, we have the 4-dimensional Chern–Weil form given by

〈Ω ∧ Ω〉 ∈ Ω4(M),

where 〈·, ·〉 denotes a suitably normalized Ad-invariant metric on g. For G =

37

Page 45: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Spin(n), this is one-half the first Pontryagin form and denoted

p1

2(P,Θ) =

−1

16π2Tr(Ω ∧ Ω);

for G = SO(n), this is the first Pontryagin form and denoted

p1(P,Θ) =−1

8π2Tr(Ω ∧ Ω);

and for G = SU(n), this is the second Chern form and denoted

c2(P,Θ) =1

8π2Tr(Ω ∧ Ω),

where Tr always denotes the trace of the standard matrix representation.

Definition 2.5.5. For G simple, the Chern–Simons 3-form of a connection Θ on a

principal G-bundle P is

α(Θ)def= 〈Θ ∧ Ω〉 − 1

6〈Θ ∧ [Θ ∧Θ]〉 ∈ Ω3(P ).

Proposition 2.5.6. The Chern–Simons 3-form has the following properties.

1. i∗xα(Θ) = −16〈θ ∧ [θ ∧ θ]〉 ∈ Ω3(G) is bi-invariant.

2. −16〈θ ∧ [θ ∧ θ]〉 ∈ H3(G; R) is the real cohomology class associated to the R-

image of the (standard) generator of H3(G; Z) ∼= Z (in the stable range of

G).

3. R∗gα(Θ) = α(Θ); i.e. α(Θ) is right-invariant.

4. dα(Θ) = 〈Ω ∧ Ω〉 ∈ Ω4(P ).

Proof. The first statement follows by i∗xΘ = θ, along with the Ad-invariance of the

Lie bracket and inner product 〈·, ·〉.

38

Page 46: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Consequently, −16〈θ ∧ [θ ∧ θ]〉 is closed and represents a de Rham cohomology

class. The form 〈θ ∧ [θ ∧ θ]〉 ∈ Λ3g∗ is, up to constant, the standard generator

for H3(g) of any simple Lie algebra. This can be seen by evaluating it on a su(2)

subalgebra. The normalization factor in the definition of 〈·, ·〉 serves to make it the

R-image of a generator of H3(G; Z) (at least in the stable range; otherwise see the

discussion below). The factor itself is unimportant for our purposes, so we merely

define it such that it satisfies this property.

The right-invariance of α(Θ) follows from the right-equivariance of Θ and Ω,

along with the Ad-invariance of trace, since

R∗gα(Θ) = 〈Adg−1Θ ∧ Adg−1Ω〉 − 1

6〈Adg−1Θ ∧ [Adg−1Θ ∧ Adg−1Θ]〉

= 〈Θ ∧ Ω〉 − 1

6〈Θ ∧ [Θ ∧Θ]〉 = α(Θ).

Finally, the derivative of α(Θ) is given by the calculation

dα(Θ) = d〈Θ ∧ Ω〉 − d1

6〈Θ ∧ [Θ ∧Θ]〉

= 〈dΘ ∧ Ω〉 − 〈Θ ∧ dΩ〉 − 1

2〈dΘ ∧ [Θ ∧Θ]〉

= 〈(Ω− 1

2[Θ ∧Θ]) ∧ Ω〉+ 〈Θ ∧ [Θ ∧ Ω]〉 − 1

2〈(Ω− 1

2[Θ ∧Θ]) ∧ [Θ ∧Θ]〉

= 〈Ω ∧ Ω〉 − 1

4〈Θ ∧ [Θ ∧ [Θ ∧Θ]]〉

= 〈Ω ∧ Ω〉.

The above equalities follow from the definition of curvature, a standard Bianchi

identity

dΩ + [Θ ∧ Ω] = 0,

the fact that 〈ψ1 ∧ [ψ2 ∧ ψ3]〉 = 〈ψ2 ∧ [ψ3 ∧ ψ1]〉, and the Jacobi identity.

We now make some remarks about the Chern–Weil and Chern–Simons forms

for G = Spin(n) with n ≤ 4, since Spin(1), Spin(2), and Spin(4) are not simple

39

Page 47: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

groups. Definitions 2.5.4 and 2.5.5 both make sense for these groups, but we should

be careful about the statement of the second property in Proposition 2.5.6. The

characteristic class p1

2(P ) ∈ H4(M ; Z) is a stable class. This means that is the

pullback of

p1

2∈ H4(BSpin; Z)

where BSpin is the direct limit lim→BSpin(n). The 3-dimensional cohomology of

Spin(n) and the 4-dimensional cohomology of BSpin(n) both stabilize at n = 5.

Thus, the inclusions Spin(5) → Spin and BSpin(5) → BSpin induce the iden-

tity maps on 3- and 4-dimensional cohomology, respectively. Therefore, the state-

ments above are clear for Spin(n) (n ≥ 5), since we have a standard generator of

H3(Spin(n); Z) ∼= Z.

The correct statement for Spin(n) (n ≤ 4) is that the cohomology class −16〈θ ∧

[θ ∧ θ]〉 ∈ H3(Spin(n); R) is the R-image of the generator of H3(Spin(5); Z) under

the inclusion Spin(n) → Spin(5). One can show Spin(4) ∼= SU(2) × SU(2) and

Spin(3) ∼= SU(2). Up to sign choices in the following notation, the generator from

H3(Spin(5); Z) ∼= Z gets sent to (1, 1) ∈ Z ⊕ Z ∼= H3(Spin(4); Z) and 2 ∈ Z ∼=

H3(Spin(3); Z) under the respective inclusions.

40

Page 48: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 3

RIGHT-INVARIANT FORMS ON A PRINCIPAL BUNDLE

In this chapter, we give an explicit description of the complex and dual complex of

right-invariant forms on P . This description will be important in the calculations of

Chapter 5. To make sense of the dual complex, we construct a Riemannian metric

on P which depends on a metric on M and a connection Θ. We also introduce

the rescaled derivative dδ, the rescaled coderivative d∗δ , and the rescaled Laplacian

Lgδ, all of which are fundamental in the spectral sequence of Chapter 4 and the

calculations of Chapter 5.

3.1 Isomorphic bi-graded cochain complex

Let G be a compact Lie group and Pπ→M a principal G-bundle with connection

Θ ∈ Ω1(P ; g). The connection picks out a horizontal subpace at each point in P ,

and hence

TP = THP ⊕ T V P,

where THP = Ker(Θ) denotes the horizontal distribution and T V P = Ker(π∗) de-

notes the vertical distribution. This induces a bi-grading on the space of differential

forms

Ωk(P ) =⊕

i+j=k

Ωi,j(P ) =⊕

i+j=k

C∞(P ; ΛiTHP ∗ ⊗ ΛjT V P ∗).

41

Page 49: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

The bi-grading notation will always be (horizontal, vertical). In other words, Ωi,j(P )

is a form which takes values on i horizontal and j vertical vectors. Furthermore, the

exterior derivative decomposes as (Lemma 3.1.1)

d = d0,1 + d1,0 + d2,−1,

where da,b : Ωi,j(P ) → Ωi+a,j+b(P ). The right action of G preserves the decompo-

sition of TP (i.e. it maps THP → THP and T V P → T V P ). It then makes sense

to talk about the space Ωi,j(P )G of right-invariant (i, j) forms. In this next section,

we show there is a natural isomorphism of bi-graded cochain complexes

Ωi(M ; Λjg∗P ), d π∗−→ Ωi,j(P )G, d.

Furthermore, the bi-graded decomposition of d in Ωi(M ; Λjg∗P ), d has a nice de-

scription. The (0, 1) component is given, up to sign, by the vector bundle homo-

morphism induced from the derivative of right-invariant vector fields on Px

Λkg∗Pdg−→ Λk+1g∗P .

The principal bundle connection Θ on P induces a vector bundle connection ∇ on

the associated bundles Λjg∗P . The (1, 0) component of d is given by the unique

extension of ∇ to the cochain complex

· · · d∇−→ Ωi(M ; Λjg∗P )d∇−→ Ωi+1(M ; Λjg∗P )

d∇−→ · · · .

Finally, the curvature is an element Ω ∈ Ω2(M ; g), and d2,−1 is given, up to sign, by

contracting along this vector-valued 2-form

Ωi(M ; Λjg∗P )ιΩ−→ Ωi+2(M ; Λj−1g∗P ).

42

Page 50: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

The bi-graded complex isomorphic to Ωi,j(P )G, d is written

......

......

0 // Ω0(M ; Λ2g∗P )

dg

OO

∇ //

ιΩ ++

Ω1(M ; Λ2g∗P )

−dg

OO

d∇ //

−ιΩ++

Ω2(M ; Λ2g∗P )

dg

OO

d∇ //

ιΩ ++

Ω3(M ; Λ2g∗P )

−dg

OO

d∇ // · · ·

0 // Ω0(M ; g∗P )

dg

OO

∇ //

ιΩ ++

Ω1(M ; g∗P )

−dg

OO

d∇ //

−ιΩ++

Ω2(M ; g∗P )

dg

OO

d∇ //

ιΩ ++

Ω3(M ; g∗P )

−dg

OO

d∇ // · · ·

0 // Ω0(M ; R)

0

OO

dM // Ω1(M ; R)

0

OO

dM // Ω2(M ; R)

0

OO

dM // Ω3(M ; R)

0

OO

dM // · · ·

0

OO

0

OO

0

OO

0

OO

We now proceed in proving the above statements.

First, let us look at the general situation of the exterior derivative on a bi-graded

complex induced from two distributions.

Lemma 3.1.1. Let P be generic manifold with two linearly independent distribu-

tions A,B ⊂ TP such that A+B = TP . This induces a bi-grading on Ω∗(P ),

Ωi,j(P ) = C∞(P,ΛiA∗ ⊗ ΛjB∗);

the exterior derivative decomposes as

d = d−1,2 + d0,1 + d1,0 + d2,−1,

where da,b : Ωi,j(P ) → Ωi+a,j+b(P ). The component d−1,2 = 0 if and only if B is

integrable, and d2,−1 = 0 if and only if A is integrable.

Proof. The bi-grading on forms is a simple consequence of

ΛkTP ∗ = Λk(A⊕B)∗ = Λk(A∗ ⊕B∗) =⊕

i+j=k

ΛiA∗ ⊗ ΛjB∗.

43

Page 51: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

To prove the second part, it suffices to show for 1-forms (due to the Leibniz rule).

Since

d : Ω0,1(P )⊕ Ω1,0(P )→⊕

i+j=2

Ωi,j(P ),

we see that d has at most 4 parts. The d−1,2 and d2,−1 components are due to the

possible non-integrality of B and A, respectively. To see this, look at the formula

for dψ where V0, V1 are vector fields:

dψ(V0, V1) = V0ψ(V1)− V1ψ(V0)− ψ([V0, V1]).

In particular, if ψ ∈ Ω1,0, and V0, V1 ∈ C∞(P ;B), then

d2,−1ψ(V0, V1) = −ψ([V0, V1]). (3.1.1)

If B is integrable, then for any such vector fields in B, [V0, V1] ∈ C∞(M ;B), and

therefore d2,−1ψ = 0. However, if B is not integrable, there exist vector fields V0, V1

such that [V0, V1]|A 6= 0. Letting ψ ∈ Ω1,0 be a differential form which is non-zero

on [V0, V1], we see that d2,−1ψ 6= 0. The same argument shows that d2,−1 = 0 if and

only if A is integrable.

Because P is a fiber bundle, the vertical distribution T V P is integrable. This

follows immediately from

π∗[V0, V1] = [π∗V0, π∗V1].

Therefore, d = d0,1 + d1,0 + d2,−1. However, THP does not have to be integrable,

and in fact, integrability of THP is equivalent to Θ being a flat connection.

We now wish to look at the sub-complex of right-invariant forms on P

Ωi,j(P )G, d → Ωi,j(P ), d.

44

Page 52: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

From (2.3.1),

T V P ∼= P × g ∼= π∗gP ,

and from (2.3.4),

THP ∼= π∗TM.

Therefore, we see that right-invariant horizontal vector fields on P are equivalent to

vector fields on M :

C∞(P, THP )G ∼= C∞(P, π∗TM)G = π∗C∞(M,TM);

and right-invariant vertical vector fields on P are equivalent to sections of the adjoint

bundle gP :

C∞(P, T V P )G ∼= C∞(P, π∗gP )G = π∗C∞(M, gP ).

We also see that right-invariant (i, j) forms on P are equivalent to i-forms on M

with values in the j-th exterior power of the dual adjoint bundle:

Ωi,j(P )G = C∞(P,ΛiTHP ∗ ⊗ ΛjT V P ∗)G ∼= C∞(P, π∗ΛiTM∗ ⊗ π∗Λjg∗P )G

= π∗C∞(M,ΛiTM∗ ⊗ Λjg∗P ) = Ωi(M ; Λig∗P ).

If V ∈ C∞(M,TM), we will denote the induced right-invariant vector field V ∈

C∞(P, THP )G (and likewise for vertical vector fields). Similarly, for ψ ∈ Ωi(M ; Λjg∗P ),

we denote the induced form ψ ∈ Ωi,j(P )G.

We see that the bi-graded complex Ωi(M ; Λjg∗P ), d is isomorphic to Ωi,j(P )G, d,

but the derivative in the first complex is so far only defined by the isomorphism. We

now wish to give a more intrinsic description of this differential. Each component

of d satisfies the Leibniz rule, so to describe d on Ωi(M ; Λjg∗P ), it suffices to say how

each component acts on Ωi(M) and C∞(M ; Λjg∗P ).

Proposition 3.1.2. For d : Ωi(M ; R) → Ωi+1(M ; R) ⊕ Ωi(M ; g∗P ), d = d1,0 and is

45

Page 53: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

the exterior derivative on M , denoted dM ; i.e. the following diagram is commutative:

Ωi(M)

π∗

dM // Ωi+1(M)

π∗⊕0

Ωi,0(P )G d1,0+d0,1// Ωi+1,0(P )G ⊕ Ωi,1(P )G

Proof. By naturality of d, we see that

dπ∗Ωi(M) = π∗dMΩi(M).

This proves commutativity of the diagram.

In fact, a converse to the above Proposition also holds.

Lemma 3.1.3. If ψ ∈ Ωi,0(P ), and d0,1ψ = 0, then ψ ∈ Ωi,0(P )G.

Proof. For ψ ∈ Ωi,0(P ), the (0, 1) component of d is given by the fiberwise derivative.

Let X1, X2, . . . , Xi be horizontal right-invariant vector fields on P , and V a vertical

vector field on P . Then, the brackets [V,Xl] = 0, and

dψ(V,X1, . . . , Xi) = V ψ(X1, . . . , Xi).

If d0,1ψ = 0, then ψ restricted to any fiber i∗xψ ∈ Ωi(Px) is a constant section in

C∞(Px, π∗ΛiTM∗). Consequently, ψ ∈ C∞(P, π∗ΛiTM∗)G.

Proposition 3.1.4. d0,1 : Ω0(M ; Λjg∗P ) → Ω0(M ; Λj+1g∗P ) is the map induced by

the vector bundle homomorphism

Λjg∗Pdg−→ Λj+1g∗P .

46

Page 54: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

In other words, the following diagram is commutative:

Ω0(M,Λjg∗P )

π∗

dg // Ω0(M,Λj+1g∗P )

π∗

Ω0,j(P )G d0,1

// Ω0,j+1(P )G

Proof. First, note that the fiberwise inclusion ix : Px → P induces an isomorphism

ix∗ : TPx∼= (T V P )x.

Therefore, at any point p ∈ P , we have the canonical isomorphism of complexes

(i∗x)p : (Λ∗T V P ∗)p≈−→ (Λ∗TP ∗

x )p.

For arbitrary ψ ∈ Ω0,j(P ), d0,1ψ ∈ Ω0,j+1(P ). This implies

(d0,1ψ)p = i∗x(d0,1ψ)p = i∗x(dψ)p = (di∗xψ)p.

The middle equality comes from i∗xd = i∗xd0,1. Therefore, for ψ = π∗ω ∈ Ω0,j(P )G,

the fiber-wise derivative (di∗x)(π∗ω) is given by (2.2.4),

Λjg∗Px

π∗

dg // Λj+1g∗Px

π∗

Ωi(Px)

G d // Ωi+1(Px)G

Therefore, at each point x ∈M the section d0,1ω is given by dgω.

Proposition 3.1.5. The component d1,0 : Ω0(M ; Λjg∗P ) → Ω1(M ; Λjg∗P ) is the co-

variant derivative ∇ induced by the connection Θ on the associated bundle Λjg∗P . In

other words, the following diagram is commutative:

Ω0(M ; Λjg∗P )

π∗

∇ // Ω1(M ; Λjg∗P )

π∗

Ω0,j(P )G d1,0

// Ω1,j(P )G

47

Page 55: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proof. Let f ∈ C∞(M ; Λjg∗P ), and

f ∈ C∞(P,Λjg∗)G ∼= Ω0,j(P )G

be the associated equivariant g-valued function. Then,

d1,0f = d1,0f

where, in the total space P , d1,0 is defined by derivating along horizontal vectors.

Given a horizontal vector field XM on M , the connection provides a lift to a vector

field XM on P , and

d1,0XM

f = dgXMf .

As defined in (2.3.5), this is covariant differentiation on the associated bundle in-

duced by the connection Θ. Therefore,

∇XMf = dgXM

f = d1,0f .

Proposition 3.1.6. d2,−1 : Ω0(M ; g∗P ) → Ω2(M ; R) is given by composing the cur-

vature with the natural evaluation; i.e. the following diagram is commutative:

Ω0(M ; g∗P )Ω //

π∗

Ω2(M ; gP ⊗ g∗P ) // Ω2(M ; R)

π∗

Ω0,1(P )G d2,−1// Ω2,0(P )G

Proof. Recalling (2.3.7), we see that the curvature Θ ∈ Ω2(M ; gP ) can be described

by taking the Lie bracket of two right-invariant horizontal vector fields and then

projecting to the vertical space. If X0, X1 are right-invariant horizontal vector fields

48

Page 56: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

on P , then by (2.3.7)

Ω(X0, X1) = −Θ([X0, X1]) ∈ C∞(P, T V P )G.

Now, let ψ ∈ Ω0(M ; g∗P ), and ψ = π∗ψ ∈ Ω0,1(P )G. As noted in (3.1.1), d2,−1 is

determined by

d2,−1ψ(X0, X1) = −ψ([X0, X1]).

Because ψ is zero on horizontal vectors, evaluating it on a vector is the same as first

projecting the vector to the vertical tangent space and then evaluating. In other

words, for any vector X in TP ,

ψ(X) = ψ(Θ(X)).

Combining the above equations, we have

d2,−1ψ(X0, X1) = −ψ(Θ([X0, X1])

)= ψ

(Ω(X0, X1)

)∈ R.

Therefore,

d2,−1ψ(X0, X1) = ψ (Ω(X0, X1))

where Ω(X0, X1) ∈ gP and ψ ∈ g∗P .

We have now described the d0,1, d1,0, d2,−1 acting on the Ωi(M ; R) and Ω0(M ; Λjg∗P ).

By using the Leibniz rule, we can extend these descriptions to arbitrary Ωi(M ; Λjg∗P ).

For the following discussion, let ω ∈ Ωi(M ; R) and η ∈ Ω0(M ; Λjg∗P ). First,

d0,1(ω ∧ η) = (d0,1ω) ∧ η + (−1)iω ∧ (d0,1η)

= (−1)iω ∧ (d0,1η).

(3.1.2)

This follows from the Leibniz rule and the fact that d0,1 : Ωi(M ; R) → Ωi(M ; g∗P )

is zero. Therefore, d0,1 : Ωi(M ; Λjg∗P ) → Ωi(M ; Λjg∗P ) is, up to sign, the vector

bundle homomorphism induced by Λjg∗Pdg→ Λj+1g∗P , and our notation becomes d0,1 =

49

Page 57: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

(−1)idg;

Ωi(M ; Λjg∗P )(−1)idg−→ Ωi(M ; Λj+1g∗P ).

Next, we see that

d1,0(ω ∧ η) = (d1,0ω) ∧ η + (−1)iω ∧ (d1,0η)

= (dMω) ∧ η + (−1)iω ∧ (∇η).

Here, we use the general symbol ∇ for the connection on each bundle Λjg∗P , as the

bundles and connections are all constructed from (P,Θ). The d0,1 satisfies

• d1,0η = ∇η for η ∈ Ω0(M ; Λjg∗P ),

• d1,0(ω ∧ η) = (dMω) ∧ η + (−1)iω ∧ (∇η).

Therefore (2.3.6), d1,0 is the standard extension of ∇ : Ω0(M ; Λjg∗P )→ Ω1(M ; Λjg∗P )

to the complex Ωi(M ; Λjg∗P ). We therefore use the notation d1,0 = d∇, and

Ω0(M ; Λjg∗P )∇−→ Ω1(M ; Λjg∗P )

d∇−→ Ω2(M ; Λjg∗P )d∇−→ · · · .

To write d2,−1 on arbitrary sections in Ω0(M ; Λjg∗P ), we observe that d2,−1 satisfies

the properties

• d2,−1η = η(Ω) for η ∈ Ω0,1,

• d2,−1(η1 ∧ η2) = (d2,−1η1) ∧ η2 + (−1)deg η1η1 ∧ (d2,−1η2).

Therefore, d2,−1 : Ω0(M ; Λjg∗P ) → Ω2(M ; Λj−1g∗P ) is simply the contraction along

the (vertical-vector)-valued 2-form Ω. We denote this by ιΩ. Evaluating on elements

of Ωi(M ; Λjg∗P ),

d2,−1(ω ∧ η) = (−1)iω ∧ (d2,−1η)

= (−1)iω ∧ (ιΩη).

50

Page 58: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

We use the notation d2,−1 = (−1)iιΩ, and

Ωi(M ; Λjg∗P )(−1)iιΩ−→ Ωi+2(M ; Λj−1g∗P ).

Combining the above descriptions of the components of d gives the following propo-

sition.

Proposition 3.1.7. The following bi-graded cochain complex is isomorphic to Ωi,j(P )G, d,

the cochain complex of right-invariant forms on P .

......

......

0 // Ω0(M ; Λ2g∗P )

dg

OO

∇ //

ιΩ ++

Ω1(M ; Λ2g∗P )

−dg

OO

d∇ //

−ιΩ++

Ω2(M ; Λ2g∗P )

dg

OO

d∇ //

ιΩ ++

Ω3(M ; Λ2g∗P )

−dg

OO

d∇ // · · ·

0 // Ω0(M ; g∗P )

dg

OO

∇ //

ιΩ ++

Ω1(M ; g∗P )

−dg

OO

d∇ //

−ιΩ++

Ω2(M ; g∗P )

dg

OO

d∇ //

ιΩ ++

Ω3(M ; g∗P )

−dg

OO

d∇ // · · ·

0 // Ω0(M ; R)

0

OO

dM // Ω1(M ; R)

0

OO

dM // Ω2(M ; R)

0

OO

dM // Ω3(M ; R)

0

OO

dM // · · ·

0

OO

0

OO

0

OO

0

OO

Remark 3.1.8. The above complex is not a commutative diagram, and d2∇ 6= 0.

Instead, we have that

0 = d2 = (d0,1 + d1,0 + d2,−1)2,

51

Page 59: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

and therefore we see the equalities

0 = (d0,1)2 = d2g

0 = d0,1d1,0 + d1,0d0,1 = ±(dgd∇ − d∇dg)

0 = d0,1d2,−1 + d1,0d1,0 + d2,−1d0,1 = dgιΩ + d2∇ + ιΩdg

0 = d1,0d2,−1 + d2,−1d1,0 = ±(d∇ιΩ − ιΩd∇)

0 = (d2,−1)2 = ι2Ω

In other words, if you start at a point in the diagram and look at another point “two

arrows away,” summing over all the possible paths between the two points gives zero.

From this point of view, the bi-graded complex resembles a spectral sequence.

3.2 The adjoint d∗ in the bi-graded complex

For Chapter 5 of this thesis, we wish to think of P as a Riemannian manifold

and study the low-dimensional harmonic forms. Given a metric g on M and a

connection Θ on P , along with a bi-invariant metric gG on G, we define a canonical

right-invariant metric gP on P . The metric gP is defined by g on the horizontal

tangent bundle THP , and gG on the vertical tangent bundle T V P . However, we are

really interested in the 1-parameter family of metrics gδ on P given by g on THP and

δ2gG on T V P . The results in Chapter 5, where we look at the limit of gδ-harmonic

forms as δ → 0, will be independent of the choice of metric gG on the fibers G.

Since each metric gδ is the pullback of metrics on TM and gP , the inner product

space of G-invariant forms on P is naturally isomorphic to the inner product space

Ωi(M ; Λjg∗P ),Λig∗ ⊗ Λjg∗G. This implies that when restricted to right-invariant

forms, the adjoint d∗gδ is given by the adjoints to dg, dM , d∇, and ιΩ.

We now proceed in more detail. Start with the data in the previous sections:

52

Page 60: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

(M,P,Θ) where Pπ→M is a principal G-bundle with connection Θ. In fact,

TP = THP ⊕ T V P = π∗(TM ⊕ gP ),

where gP → M is the adjoint bundle P ×Ad g. We introduce the choice of a Rie-

mannian metric g on M (which is a metric on the bundle TM). This is denoted

(M,P, g,Θ). Next, we must choose an Ad-invariant metric gG on g. Since G is

compact, such metrics always exist. In fact, there is a 1-1 correspondence

Bi-invariant metrics on G ←→ Ad-invariant metrics on g,

and we will sometimes blur the distinction between the two. Since gG is Ad-invariant,

this induces a metric on the adjoint bundle gP →M , which we also denote gG.

We now have a natural right-invariant metric gP on P given by

gPdef= π∗(g ⊕ gG).

In other words, the metric on the horizontal spaces is defined by lifting the metric

from the base to a metric on THP , the metric on the fibers is given by the metric

on G, and the horizontal and vertical subspaces are defined to be perpendicular.

Calculating at a point, this looks like

(TP, gP )def= (THP, g)⊕ (T V P, gG)

gP (v1, v2) = g(π∗v1, π∗v2) + gG(Θ(v1),Θ(v2)).

The metric on gP induces metrics on all the spaces of forms

(Ωi,j(P ), π∗(Λig∗ ⊗ Λjg∗G)

).

Restricting to the subcomplex of right-invariant forms, we have the following iso-

53

Page 61: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

morphisms of cochain complexes with metric:

(Ωi,j(P ), π∗(Λig∗ ⊗ Λjg∗G)

)G ∼=(C∞(P, π∗(ΛiTM∗ ⊗ Λjg∗P )), π∗(Λig∗ ⊗ Λjg∗G)

)G

= π∗(C∞(M,ΛiTM∗ ⊗ Λjg∗P ),Λig∗ ⊗ Λjg∗G

)= π∗

(Ωi(M ; Λjg∗P ),Λig∗ ⊗ Λjg∗G

).

All of the metrics are induced from the metrics g and gG, and we will always think

of gG as fixed, while g can be any element in Met(M).

As defined in (2.4.1), let

d∗gP : Ωk(P )→ Ωk−1(P )

be the adjoint to d on Ω∗(P ), and let d∗ denote the adjoint to the differential d on

Ωi(M ; Λjg∗P ). Then, using the inclusion of cochain complexes with metric

Ωi,j(P )G, d → Ωi,j(P ), d,

we have the induced inclusion on the dual complexes

(Ωi(M ; Λjg∗P ), d∗

) ∼= (Ωi,j(P )G, d∗gP

)→

(Ωi,j(P ), d∗gP

). (3.2.1)

Under the bi-grading,

d∗ = d0,1∗ + d1,0∗ + d2,−1∗,

where

da,b∗ : Ωi(M ; Λjg∗P )→ Ωi−a(M ; Λj−bg∗P ).

Using our description of the differentials da,b in Proposition 3.1.7, we write this as

d0,1∗ = (−1)i(dg)∗, d1,0∗ = (d∇)∗, d2,−1∗ = (−1)i(ιΩ)∗.

where the adjoints are induced by the metrics g and gG on TM and gP , respectively.

54

Page 62: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

We therefore have the following complex on right-invariant forms determined by

the coderivative associated to the right-invariant metric gP .

...

d∗g

...

−d∗g

...

d∗g

...

−d∗g

0 Ω0(M ; Λ2g∗P )oo

d∗g

Ω1(M ; Λ2g∗P )d∗∇oo

−d∗g

Ω2(M ; Λ2g∗P )d∗∇oo

d∗g

ι∗Ω

kk

Ω3(M ; Λ2g∗P )d∗∇oo

−d∗g

−ι∗Ω

kk

· · ·d∗∇oo

0 Ω0(M ; g∗P )oo

0

Ω1(M ; g∗P )d∗∇oo

0

Ω2(M ; g∗P )d∗∇oo

0

ι∗Ω

kk

Ω3(M ; g∗P )d∗∇oo

0

−ι∗Ω

kk

· · ·d∗∇oo

ι∗Ω

kk

0 Ω0(M ; R)oo

Ω1(M ; R)d∗Moo

Ω2(M ; R)d∗Moo

ι∗Ω

kk

Ω3(M ; R)d∗Moo

−ι∗Ω

kk

· · ·d∗Moo

ι∗Ω

kk

0 0 0 0

In particular, the adjoint d∗M : Ωi(M ; R) → Ωi−1(M ; R) is the usual adjoint with

respect to the metric g, and the adjoint d∗g is induced from the usual adjoint to the

Lie algebra derivative with respect to the metric gG.

3.3 Rescaled derivatives dδ and d∗δ

In general, calculating the harmonic forms on P is still very difficult. It turns

out that if we shrink the metric on the fiber while keeping the metric on the base

the same, the harmonic forms become easier to compute and have an interesting

description. This limiting process is referred to as the adiabatic limit. We now

define the 1-parameter family of metrics gδ, the rescaled derivative dδ, the rescaled

coderivative d∗δ , and the rescaled Laplacian Lgδwhich is an operator on the fixed

inner product space

Lpgδ

: (Ωp(P ), gP )→ (Ωp(P ), gP ) .

Remark 3.3.1. The standard notation in the literature is gδ = δ−2g ⊕ gG with

δ → 0, or expanding the base relative to the fibers. Since harmonic forms remain

55

Page 63: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

invariant under a global rescaling, this is equivalent to shrinking the fiber relative

to the base. The notation used from this point on will coincide with the literature,

but the author finds it conceptually more convenient to think of this as shrinking the

fiber.

As before, we choose an Ad-invariant metric gG on g (or a bi-invariant metric

on the fiber G). Just as we defined a metric gP = π∗(g ⊕ gG), define a 1-parameter

family of metrics on P by

gδdef= π∗(δ−2g ⊕ gG). (3.3.1)

For any δ > 0, gδ is a right-invariant Riemannian metric on P . Because the inner

product space Ω∗(P, gδ) is changing for each δ, we have a 1-parameter family of inner

product spaces and operators. It is therefore convenient to fix the inner product by

introducing the isometry

ρδ : (Ωi,j(P ), gδ)→ (Ωi,j(P ), gP )

φ 7→ δiφ

Conjugating by this isometry gives a 1-paramter family of operators on the fixed

inner product space (Ω∗(P ), gP ). Let ∆gδbe the usual Hodge Laplacian

∆pgδ

def= dd∗gδ + d∗gδd : (Ωp(P ), gδ)→ (Ωp(P ), gδ),

and define the operators dδ, d∗δ , and Lδ on the space (Ω∗(P ), gP ) by

dδdef= ρδdρ

−1δ

d∗δdef= ρδd

∗gδρ−1δ

Lgδ

def= ρδ∆gδ

ρ−1δ = dδd

∗δ + d∗δdδ

56

Page 64: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

In general, rescaling a metric by a global factor δ2 introduces powers of δ on the

∗ operator (see Lemma 2.4.5). Therefore, while d remains invariant, d∗ picks up

powers of δ. In our situation, combined with the bi-grading, this looks like

d = d0,1 + d1,0 + d2,−1

d∗gδ = d0,1∗ + δ2d1,0∗ + δ4d2,−1∗

where ∗ denotes the adjoint with respect to the original metric. Since the conjugation

action also introduces powers of δ, then the components of both dδ and d∗δ pick up

powers of δ according to the bi-grading:

dδ = d0,1 + δd1,0 + δ2d2,−1

d∗δ = d0,1 + δd1,0∗ + δ2d2,−1∗

When applied to the complex of right-invariant forms, which is isomorphic to Ωi(M ; Λjg∗P )

(as described in Proposition 3.1.7 and (3.3.3))

dδ = ±dg + δd∇ ± δ2ιΩ

d∗δ = ±d∗g + δd∗∇ ± δ2ι∗Ω

57

Page 65: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Therefore, Ωi,j(P )G, dδ is isomorphic to the cochain complex

......

......

0 // Ω0(M ; Λ2g∗P )

dg

OO

δ∇ //

δ2ιΩ++

Ω1(M ; Λ2g∗P )

−dg

OO

δd∇ //

−δ2ιΩ++

Ω2(M ; Λ2g∗P )

dg

OO

δd∇ //

δ2ιΩ++

Ω3(M ; Λ2g∗P )

−dg

OO

δd∇ // · · ·

0 // Ω0(M ; g∗P )

dg

OO

δ∇ //

δ2ιΩ++

Ω1(M ; g∗P )

−dg

OO

δd∇ //

−δ2ιΩ++

Ω2(M ; g∗P )

dg

OO

δd∇ //

δ2ιΩ++

Ω3(M ; g∗P )

−dg

OO

δd∇ // · · ·

0 // Ω0(M ; R)

0

OO

δdM // Ω1(M ; R)

0

OO

δdM // Ω2(M ; R)

0

OO

δdM // Ω3(M ; R)

0

OO

δdM // · · ·

0

OO

0

OO

0

OO

0

OO

(3.3.2)

Likewise, the dual complex Ωi,j(P )G, d∗δ is isomorphic to

...

d∗g

...

−d∗g

...

d∗g

...

−d∗g

0 Ω0(M ; Λ2g∗P )oo

d∗g

Ω1(M ; Λ2g∗P )δd∗∇oo

−d∗g

Ω2(M ; Λ2g∗P )δd∗∇oo

d∗g

δ2ι∗Ω

kk

Ω3(M ; Λ2g∗P )δd∗∇oo

−d∗g

−δ2ι∗Ω

kk

· · ·δd∗∇oo

0 Ω0(M ; g∗P )oo

0

Ω1(M ; g∗P )δd∗∇oo

0

Ω2(M ; g∗P )δd∗∇oo

0

δ2ι∗Ω

kk

Ω3(M ; g∗P )δd∗∇oo

0

−δ2ι∗Ω

kk

· · ·δd∗∇oo

δ2ι∗Ω

kk

0 Ω0(M ; R)oo

Ω1(M ; R)δd∗Moo

Ω2(M ; R)δd∗Moo

δ2ι∗Ω

kk

Ω3(M ; R)δd∗Moo

−δ2ι∗Ω

kk

· · ·δd∗Moo

δ2ι∗Ω

kk

0 0 0 0(3.3.3)

These two complexes will be very important in the calculations of Chapter 5. For

the reader’s convenience, these two pictures are reproduced on an extra page at the

end of the thesis.

58

Page 66: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 4

ADIABATIC SPECTRAL SEQUENCE

Our goal in the next chapter is to calculate, for k = 1, 2, 3, the harmonic k-forms

on P in the adiabatic limit. In other words, if ∆kgδ

: (Ωk(P ), gδ) → (Ωk(P ), gδ) is

the Hodge Laplacian, then what does

limδ→0

Ker ∆kgδ

look like? The most obvious question is whether this even makes sense. In fact, there

is a smooth extension of Ker ∆gδto δ = 0 ([MM]). The more important question

for us is how to calculate this limit. This comes from interpreting limδ→0 Ker ∆gδin

terms of a spectral sequence. Roughly, given a form ω ∈ Ωp(P ), we say that ω ∈ EpK

if there is a 1-parameter family of forms ω(δ) on the closed interval [0, 1] such that

ω(0) = ω, and both dδω(δ) and d∗δω(δ) vanish to order K as δ goes to 0. In other

words, if ω can be extended to a family of forms that is dδ-closed and d∗δ-coclosed

up to order K at δ = 0, then ω ∈ EpK . This filtration on the space of forms is then

isomorphic to the Leray–Serre spectral sequence.

This idea was developed by Mazzeo and Melrose in [MM] and expanded in [Dai]

and [For]. In particular, Forman made explicit the spectral sequence structure,

including the differentials. The following is a summary of his treatment. The

spectral sequence structure holds for more general situations than ours, so we start

with a more general discussion, and add in extra hypotheses as needed. In particular,

59

Page 67: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

all statements will hold for our situation of a principal bundle with metric given by

(M,P, g,Θ).

Remark 4.0.2. Warning! In Forman’s paper, his bi-grading notation is opposite

the notation here. Whereas our notation is always (horizontal, vertical), Forman’s

notation is (vertical, horizontal). This also means that the distributions A and B

are reversed. In the specific example we look at, the bi-grading naturally shows up

in the notation Ωi(M ; Λjg∗P ), and it seems more natural to call this space Ωi,j(P )

than the opposite.

4.1 Leray–Serre spectral sequence

In general, a cohomology spectral sequence is comprised of

• a sequence of “pages” or cochain complexes Ep0 , E

p1 , E

p2 , · · · , Ep

∞. Frequently,

these are equipped with a bi-grading EpK =

⊕i+j=p

Ei,jK .

• differentials dK : EpK → Ep+1

K such that Hp(E∗K , dK) = Ep

K+1. If there is a

bi-grading present, then dK : Ei,jK → Ei+K,j−K+1

K .

Normally, there are isomorphisms relating the E2 and E∞ pages to something with

a more intrinsic description. The example we deal with is the Leray–Serre spectral

sequence associated to a fibration of compact manifolds F → Pπ→M . Then,

• Ei,j2∼= H i(M ;Hj(F ;H)) (if π1(M) acts trivially on P ),

• Ep∞ gives Hp(M ;H) in the sense that En,p−n

∞ n is isomorphic to the quotients

F pn/F

pn+1 of a filtration 0 ⊂ F p

p ⊂ · · · ⊂ F p0 = Hp(P ;H).

In other words, we have a method of relating the cohomology of the total space

of a fibration with the cohomology of its base and fiber. This allows for powerful

calculations using very formal methods. For example, . . .

60

Page 68: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Example 4.1.1. Let’s apply the Leray–Serre spectral sequence for integral cohomol-

ogy to the universal principal G-bundle EG → BG, where G is compact, simple,

and simply-connected. Since H1(G; Z) = H2(G; Z) = 0, and H3(G; Z) ∼= Z, then,

the E2 through E4 pages look like the following, with the only non-trivial differential

being d4 on the E4 page.

H3(G; Z)

d4

**TTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTTT· · ·

0 0 0 0 0

0 0 0 0 0

H0(BG; Z) H1(BG; Z) H2(BG; Z) H3(BG; Z) H4(BG; Z)

EG is a contractible space, and therefore the E∞ page must be zero (except for

E0,0∞∼= H0(EG; Z) ∼= Z). Consequently, H i(BG; Z) = 0 for 1 ≤ i ≤ 3, and

d4 : H3(G; Z)≈−→ H4(BG; Z).

The generator of H3(G; Z) gets sent to a universal 4-dimensional characteristic

class. For G = SU(n) (n ≥ 2) this is the c2 class, and for G = Spin(n) (n ≥ 5)

this is the p1

2class.

4.2 Adiabatic spectral sequence

The following spectral sequence is defined in a very general situation, and the

treatment is taken almost directly from Section 0 of [For]. Once we impose certain

geometric conditions, this gives analytic information. Suppose we have a compact

61

Page 69: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Riemannian manifold (P, gP ) (not necessarily a principal bundle) with two orthog-

onal distributions A and B in TP such that

A⊕B = TP

and gP = gA⊕gB. Neither A nor B need to be integrable. This induces a bi-grading

on the space of differential forms

Ωp(P ) =⊕

i+j=p

Ωi,j(P ), Ωi,j(P ) = C∞(P,ΛiA∗ ⊕ ΛjB∗);

the exterior derivative and its dual decompose as

d = d−1,2 + d0,1 + d1,0 + d2,−1, da,b : Ωi,j(P )→ Ωi+a,j+b(P ),

d∗g = (d−1,2)∗ + (d0,1)∗ + (d1,0)∗ + (d2,−1)∗, da,b∗ : Ωi,j(P )→ Ωi−a,j−b(P ).

Definition 4.2.1. Define the 1-parameter family of metrics gδ by

gδdef= δ−2gA ⊕ gB.

Introducing the isometry (for any δ > 0)

ρδ : (Ωi,j, gδ)→ (Ωi,j, gP )

φ 7→ δiφ

gives rise to operators dδ, d∗δ , and Lgδ

on the space (Ω∗(P ), gP ) defined by

dδdef= ρδdρ

−1δ , dδ = δ−1d−1,2 + d0,1 + δd1,0 + δ2d2,−1, (4.2.1)

d∗δdef= ρδd

∗gδρ−1δ , d∗δ = δ−1d−1,2∗ + d0,1∗ + δd1,0∗ + δ2d2,−1∗, (4.2.2)

Lgδ

def= ρδ∆gδ

ρ−1δ = dδd

∗δ + d∗δdδ. (4.2.3)

62

Page 70: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

In addition to fixing the Hilbert space, the isometry ρδ produces the above factors

of δ, which naturally give the spectral sequence structure. First, we define the

terms Ei,jK and then give the differentials and show why this is a spectral sequence.

This is the opposite of what normally occurs, since a spectral sequence is usually

defined inductively. In dealing with Ei,jK as a spectral sequence, it is useful to

treat δ as a formal variable. However, from a geometric point of view, δ is the

value of a parameter. Note that these two points of view agree on polynomials in

Ω∗(P )[δ]. In other words, dδ and d∗δ act formally on Ω∗(P )[δ], but if we evaluate

at some δ = 0, then this agrees with the actual dδ and d∗δ acting on Ω∗(P ). This

is most conveniently expressed in the commutative diagram, where evδ denotes the

evaluation map at some δ > 0.

Ω∗(P )[δ]dδ //

evδ

Ω∗(P )[δ]

evδ

Ω∗(P )

dδ // Ω∗(P )

We will also deal with infinite formal powers series, and we will then be careful to

make the distinction between Lgδdefined in (4.2.3) and the formal

Lδ : Ω∗(P )[|δ|]→ Ω∗(P )[|δ|] (4.2.4)

given by formally applying d∗δdδ + dδd∗δ .

Definition 4.2.2.

Ei,jK

def= ω ∈ Ωi,j(P ) | ∃ω1, . . . , ωl ∈ Ωi+j(P ) with

dδ(ω + δω1 + · · ·+ δlωl) ∈ δKΩ∗(P )[δ]

d∗δ(ω + δω1 + · · ·+ δlωl) ∈ δKΩ∗(P )[δ]

The above definition is very formal, and hence useful for calculations. However,

63

Page 71: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

we could think of this in a more geometric way. Let Ωi+j(P )→ [0, 1] be the trivial

(infinite-dimensional) vector bundle over the interval [0, 1]. The polynomial ω +

δω1 + · · ·+ δlωl can be thought of as the Taylor approximation at δ = 0 of a section

ω ∈ C∞([0, 1],Ωi+j(P )).

Since for any δ > 0 we have the operators dδ and d∗δ , we can give an equivalent

definition

Ei,jK =ω ∈ Ωi,j(P ) | ∃ω ∈ C∞([0, 1],Ωi+j(P )) such that ω(0) = ω

and as δ → 0, dδω = 0 +O(δK); d∗δω = 0 +O(δK)

By definition, Ωi,j(P ) = Ei,j−1 ⊇ Ei,j

0 · · · ⊇ Ei,jK ⊇ Ei,j

K+1 ⊇ · · · ⊇ Ei,j∞ . Thus,

we have a filtration on Ωi,j(P ). (The E−1 term is included, since dδ and d∗δ have a

negative power of δ if the distribution B is not integrable.) The space Ei,j∞ is defined

by

Ei,j∞

def=

⋂K

Ei,jK ,

and can be described by forms ω ∈ Ωi,j(P ) that, when viewed as a boundary value

at δ = 0 in the space of sections C∞([0, 1],Ωi+j(P )), can be extended so that dδ and

d∗δ of the extension vanish up to any order of δ.

We now focus on the spectral sequence structure, and consequently we deal with

δ formally. To define the differentials which make Ei,jK into a spectral sequence,

first let

πK : Ω∗(P )→ EK

denote the orthogonal projection using the metric g. Suppose ω ∈ EK . Then, we

know there exists a (non-unique) extension ωδ = ω+δω1+· · · such that dδωδ, d∗δωδ ∈

64

Page 72: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

δKΩ∗(P )[δ]. To any extension ωδ, define

dKωδdef= lim

δ→0δ−Kdδωδ = lim

δ→0δ−Kdδ(ω + δω1 + δ2ω2 + · · · ).

Though this depends on the choice of extension,

πKdKπK : EK → EK

is well-defined. The same thing can also be done to define

πKd∗KπK : EK → EK .

We then form the second order operator

MKdef= (πKdKπK) (πKd

∗KπK) + (πKd

∗KπK) (πKdKπK) .

Theorem 4.2.3. (Theorem 2.5 [For])

1. (πKdKπK)2 = (πKd∗KπK)2 = 0.

2. Ker MK= Ker(πKdkπK) ∩Ker(πKd∗KπK) = EK+1.

3. πKdKπK : Ei,jK → Ei+K,j−K+1

K .

The maps πKdKπK form differentials in the complex E•K . However, instead of

taking the ordinary cohomology and dealing with equivalences classes of cocycles,

we take the Hodge cohomology of the complex E•K . Therefore, E•

K+1 ⊂ Ω∗(P );

the cochains of the new complex are still represented by differential forms, and the

sequence of complexes Ei,jK , πKdKπK form a spectral sequence.

Again, this spectral sequence is defined for arbitrary orthogonal distributions A

and B such that A⊕B = TP . In general, one can describe the convergence of this

sequence using the language of Laurent cohomology. Since this is not important for

65

Page 73: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

our later discussion, we don’t go into details. Roughly speaking, though, the Ep∞

term gives a basis for KerLδ acting formally on Laurent series with differential forms

as coefficients. This is then isomorphic to formal Laurent series with coefficients in

Hp(P ). In particular, dimEp∞ = dimHp(P ), where Ep

∞ is a finite-dimensional

subspace of differential forms. However, when B is integrable, we have the following

isomorphism.

Proposition 4.2.4. (Corollary 4.4 [For]) If the distribution B is integrable, then the

spectral sequence Ei,jK , πKdKπK is isomorphic to the standard Leray–Serre spectral

sequence for the foliation associated to B.

Before discussing the relationship with harmonic forms in the adiabatic limit,

we prove an important (for our purposes) lemma that follows from the above state-

ments.

Proposition 4.2.5. Let N = N(p) denote the page where the portion of the spectral

sequence calculating Hp(P ) collapses. i.e. N is such that

EpN(p)−1 6= Ep

N(p) = · · · = Ep∞.

Suppose

ωδ = v + δω1 + δ2ω2 + · · ·

with

ωl ∈ E⊥∞ (1 ≤ l ≤ L)

and

dδωδ, d∗δωδ ∈ δN+LΩ∗(P )[δ].

Then, for l ≤ L, the terms ωl are unique.

Proof. Suppose v + δω1 + · · · and v + δφ1 + · · · are such that ωl, φl ⊥ Ep∞ for all

66

Page 74: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

l ≥ 1. Then, we have that

dδ(v + δω1 + δ2ω2 + · · · )− dδ(v + δφ1 + δ2φ2 + · · · )

= dδ(δ(ω1 − φ1) + δ2(ω2 − φ2) + · · · ) ∈ δN+LΩ∗(P )[δ]

d∗δ(v + δω1 + δ2ω2 + · · · )− d∗δ(v + δφ1 + δ2φ2 + · · · )

= d∗δ(δ(ω1 − φ1) + δ2(ω2 − φ2) + · · · ) ∈ δN+LΩ∗(P )[δ]

Hence, ω1 − φ1 ∈ EpN+L−1 = Ep

∞, but we know that ω1 − φ1 ∈ (Ep∞)⊥. Therefore,

ω1 = φ1, and we continue inductively to show uniqueness of the higher order terms

for l ≤ L.

Theorem 4.2.6. (Theorem 3.1 [For]) Suppose ω ∈ Ei,j∞ . Then there is a unique

formal power series

ωδ = ω + δω1 + δ2ω2 + · · · ∈ Ωi+j(P )[|δ|]

such that

ωl ⊥ E∞ ∀ l ≥ 1,

and formally

dδωδ = d∗δωδ = 0.

Proof. The existence of a formally harmonic power series for ω ∈ Ei,j∞ follows from

the definition of E∞. The uniqueness of such a power series, assuming the higher

order terms in E⊥∞, follows by Proposition 4.2.5 above.

These formal power series will be useful, so we give them a seperate notation.

Definition 4.2.7.

Ei,j∞,δ

def= ωδ ∈ Ω∗(P )[|δ|] | ω ∈ Ωi,j(P ) ; ωl ∈ E⊥

∞ (l ≥ 1)

dδωδ = d∗δωδ = 0

67

Page 75: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Note that there is a bijective map

Ei,j∞ ←→ Ei,j

∞,δ,

and the difference between the two R-vector spaces is that Ei,j∞ ⊂ Ωi,j(P ), while

Ei,j∞,δ ⊂ Ω∗(P )[|δ|]. We also note that

KerLkδ = ωδ ∈ Ωk(P )[|δ|] | dδωδ = d∗δωδ = 0.

Corollary 4.2.8. The space Ek∞,δ[|δ|] = Ek

∞,δ ⊗R R[|δ|], considered as a finite-

dimensional R[|δ|]-module, equals the space of formally Lδ-harmonic power series:

Ek∞,δ[|δ|] = KerLk

δ ⊂ Ωk(P )[|δ|].

4.3 Relation to harmonic forms

To discuss the convergence of EpK , πKdKπK in relation to harmonic forms on

P , we must impose further conditions. The “vertical” distribution B must be inte-

grable, and the metric gP on P must be “bundle-like.” This amounts to saying that

(P, gP ) with distribution B locally has the structure of a Riemannian submersion

(see (H1) in Section 5 of [For] for more specifics). In particular, if

F → P →M

is a fiber bundle of compact manifolds, with P → M a Riemannian submersion,

then the metric and vertical distribution satisfy the necessary conditions. Also, if

G is a compact Lie group with bi-invariant metric acting on (P, gP ) such that the

orbits all have the same dimension, then the foliation of P associated to the orbits

of G satisfies the hypotheses. The case of a principal bundle, which is our focus, is

an example of both of these. All further discussion assumes these hypotheses are

68

Page 76: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

satisfied.

We now wish to discuss the convergence of the spectral sequence in terms of the

adiabatic limit of harmonic forms. Above, we treated δ as a formal variable. Now,

we wish to also think of it as a number in [0, 1] parameterizing the operators. To do

so, we should be careful to distinguish between the analytically defined Laplacians

(∆gδand Lgδ

) and the formal Laplacian Lδ. As a reminder,

∆kgδ

: (Ωk(P ), gδ)→ (Ωk(P ), gδ) ∀δ ∈ (0, 1]

is the 1-parameter family of standard Hodge Laplacians with respect to the metric

gδ, and

Lkgδ

: (Ωk(P ), g)→ (Ωk(P ), g) ∀δ ∈ (0, 1]

is the 1-parameter family of rescaled Laplacian defined by (4.2.3)

ρδ∆kgδρ−1

δ = Lkgδ.

The formal operator on power series is denoted

Lkδ : Ωk(P )[|δ|]→ Ωk(P )[|δ|].

We then define the spaces Ker ∆k0 and KerLk

0 by

Ker ∆k0 = lim

δ→0Ker ∆k

gδ= ω ∈ Ωk(P ) | ∃ω ∈ C∞([0, 1],Ωk(P ));

ω(0) = ω; ∆kgδω(δ) = 0 ∀δ > 0

(4.3.1)

and

KerLk0 = lim

δ→0KerLk

gδ= ω ∈ Ωk(P ) | ∃ω ∈ C∞([0, 1],Ωk(P ));

ω(0) = ω; Lkgδω(δ) = 0 ∀δ > 0

(4.3.2)

In other words, an element ω ∈ KerLk0 if and only if it is the limit of a 1-parameter

69

Page 77: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

family of Lgδ-harmonic forms. If ω is a smooth section of Ωk(P ) → [0, 1] that is

Lkgδ

-harmonic for all δ > 0, then it follows that the associated Taylor series at δ = 0,

which we denote ωδ ∈ Ωk(P )[|δ|], is formally Lkδ harmonic. In other words,

ω ∼δ=0

ωδ = ω + δω1 + δ2ω2 + · · ·

and

Lδ(ω + δω + δ2ω2 + · · · ) = 0.

This gives a map KerLk0 → Ek

∞. In fact, this map is an equality.

Theorem 4.3.1. (Corollary 5.18 and Theorem 5.21 [For]) The Lgδ-harmonic forms

converge to E∞ as δ → 0; i.e.

limδ→0

KerLkgδ

= limδ→0

ρδ Ker ∆δ = Ek∞ ⊂ Ωk(P ).

In fact, the finite-dimensional vector spaces KerLgδextend smoothly to δ = 0.

More specifically, the spaces KerLpgδ

form a C∞ map from [0,1] to the space of

(dimHk(P ))-dimensional subspaces of k-forms on P .

Corollary 4.3.2. (Corollary 5.22 [For], Corollary 18 [MM]) The finite-dimensional

vector spaces Ker ∆kgδ

extend smoothly to δ = 0.

Therefore, in addition to the spaces KerLk0 and Ker ∆k

0 being well-defined, the

smoothness property allows us to discuss the 1-parameter families

C∞([0, 1],KerLkgδ

) and C∞([0, 1],Ker ∆kgδ

).

This also gives the notion of the Taylor series at δ = 0 of a family of harmonic forms.

To see the relationship of KerL0 with Ker ∆0, remember that for δ > 0

ρδ∆gδρ−1

δ = Lgδ, Ker ∆gδ

= ρ−1δ KerLgδ

.

70

Page 78: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Since the isometry ρ−1δ involves dividing by powers of δ, we must be careful when

applying at δ = 0. To avoid these problems at δ = 0, we will talk about the subspace

of power series in Ω∗(P )[|δ|] that are in the image of ρδ. In other words, it makes

sense to apply ρ−1δ on them. We denote this ρδΩ

p(P )[|δ|].

Proposition 4.3.3. Suppose there is a power series of the form

ω + δω1 + · · ·+ δkωk +O(δk+1) ∈ KerLkδ ,

and ρ−1δ is defined on this power series. Then, applying ρ−1

δ and taking the constant

term gives an element of Ker ∆k0:

(ρ−1

δ (ω + δω1 + · · ·+ δkωk))

δ=0=

= ω0,k + ω1,k−11 + ω2,k−2

2 + · · ·+ ωk,0k ∈ Ker ∆k

0 ⊂ Ωk(P ).

Proof. Due to Theorem 4.3.1, there exists a family of Lgδ-harmonic forms

ω ∈ C∞([0, 1],KerLgδ)

such that, close to δ = 0,

ω(δ) ∼δ=0

ω + δω1 + · · ·+ δkωk +O(δk+1).

Under the isometry ρ−1δ ,

ρ−1δ ω ∈ C∞([0, 1],Ker ∆k

gδ),

and

ρ−1δ ω ∼

δ=0ρ−1

δ

(ω + δω1 + · · ·+ δkωk +O(δk+1)

).

The isometry ρ−1δ divides at most by δk. Therefore,

ρ−1δ

(ω + δω1 + · · ·+ δkωk +O(δk+1)

)=

(ω0,k + ω1,k−1

1 + ω2,k−22 + · · ·+ ωk,0

k

)+O(δ).

71

Page 79: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Consequently,

(ρ−1

δ ω)(0) =

(ρ−1

δ (ω + δω1 + · · ·+ δkωk))

δ=0

= ω0,k + ω1,k−11 + ω2,k−2

2 + · · ·+ ωk,0k ∈ Ker ∆k

0.

Corollary 4.3.4. There is an inclusion of vector spaces (as subspaces of Ω∗(P ))

Ei,0∞ ⊂ Ker ∆i

0 ⊂ Ωi(P ).

Proof. Let ω ∈ Ei,0∞ . By definition, there exists a power series of the form

ω +O(δ) ∈ KerLδ,

and consequently

δkω +O(δk+1) ∈ KerLδ.

Applying Proposition 4.3.3, we see that

(ρ−1

δ (δkω))

δ=0= ω ∈ Ker ∆k

0.

Proposition 4.3.5. If ωδ ∈ ρδΩk(P )[δ] is a finite polynomial, and

ωδ ∈ KerLδ,

then for any δ ≥ 0,

ρ−1δ ωδ ∈ Ker ∆gδ

.

Proof. For any δ > 0, ωδ ∈ KerLgδ, since Lδ = Lgδ

on polynomials. For all δ > 0,

72

Page 80: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

we have the equality

ρ−1δ KerLgδ

= Ker ∆gδ,

and therefore ρ−1δ ωδ ∈ Ker ∆gδ

for δ > 0. If ρ−1δ is defined on ωδ at δ = 0, then

(ρ−1δ ωδ)δ=0 ∈ Ker ∆0.

For us, the usefulness comes from combining Proposition 4.2.5 and Proposition

4.3.3, along with the isomorphism of EK with the Leray–Serre spectral sequence.

First, we wish to calculate the adiabatic limit of Lgδ-harmonic forms. Given a

differential form, or a polynomial of forms ωδ ∈ Ωp(P )[δ] (or 1-parameter family of

forms), it is reasonable to calculate that dδ and d∗δ are 0 up to some power of δ.

Thus, one can often show that

limδ→0

Lgδωδ = 0,

and even that it is 0 up some order of δ. A priori, that does not imply that ωδ(0)

is the limit of some Lgδ-harmonic form. But, the spectral sequence interpretation

tells us that if it vanishes to a certain order, i.e.

limδ→0

dδωδ = 0(δN), limδ→0

d∗δωδ = 0(δN),

then, higher order terms can be added so that it vanishes to any order of δ. Specif-

ically, we only need to worry about making it harmonic up to an order determined

by when the Leray–Serre spectral sequence collapses. This also applies to obtaining

information about higher order terms in the Taylor series. These higher order terms

can become important under the ρ−1δ isometry. Finally, Proposition 4.3.3 allows us

to describe the adiabatic limit of ∆gδharmonic forms in terms of the Lgδ

harmonic

forms.

73

Page 81: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 5

HARMONIC FORMS ON A PRINCIPAL BUNDLE IN THE ADIABATIC LIMIT

Let (M,P, g,Θ) denote the setup of a principal G-bundle P with connection Θ

over a Riemannian manifold (M, g), as described in Chapter 3. Let gG be any Ad-

invariant metric on g, (or equivalently a bi-invariant metric on G). The spectral se-

quence machinery from Chapter 4 makes it possible to calculate the low-dimensional

harmonic forms on P in the adiabatic limit. Given a topological assumption about

P , namely that its first real characteristic class is 0, we then see the Chern–Simons

1- and 3-forms in the Hodge cohomology. In the following calculations, we use both

the machinery of Chapter 4 and the complexes described in Chapter 3. An extra

sheet with both complexes has been attached at the end of this document for the

reader’s convenience. In the following order, we calculate:

• Ker ∆1 ⊂ Ω1(P ) when G = U(n),

• Ker ∆1 ⊂ Ω1(P ) and Ker ∆20 ⊂ Ω2(P ) when G is semisimple,

• Ker ∆30 ⊂ Ω3(P ) when G is simple.

While the main goal is the calculation of Ker ∆30 for G simple, the other simpler

cases are calculated in an analogous fashion, providing a good warm-up for the

main proof.

74

Page 82: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

5.1 1-forms when G = U(n)

We now examine the case where G = U(n) (n ≥ 1). Assume we have (M,P, g,Θ),

and let gG be any Ad-invariant metric on g. This produces the 1-parameter family

of metrics gδ as described in Chapter 3.

The part of the Leray–Serre spectral sequence which calculates the 1-dimensional

real cohomology is

H1(U(n); R)d2

++WWWWWWWWWWWWWWWWWWWWWW· · ·

H0(M ; R) H1(M ; R) H2(M ; R)

The space H1(U(n); R) is canonically isomorphic to R, and the generator is sent to

c1(P ) by the d2 differential. If c1(P ) = 0 ∈ H2(M ; R), then the d2 differential is 0,

and the portion of spectral sequence calculating H1(P ; R) collapses at N = N(1) =

2. Therefore,

E0,1∞∼= H1(U(n); R) ∼= R, E1,0

∞∼= H1(M ; R). (5.1.1)

If c1(P ) 6= 0, then d2 is injective on H1(U(n); R). Therefore,

E0,1∞ = Ker d2

(H1(U(n); R)

)= 0, E1,0

∞ = E1,02∼= H1(M ; R). (5.1.2)

Remark 5.1.1. In this situation, we can calculate the harmonic forms explicitly,

making the spectral sequence machinery unnecessary. However, we still use the

interpretation of d, d∗ in terms of the bi-graded complexes (3.3.2) and (3.3.3).

In the following, i2πTr(Θ) ∈ Ω1(P ) is the Chern–Simons 1-form for a uni-

tary bundle (Definition 2.5.2), and H1g(M) is the finite-dimensional vector space

75

Page 83: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

of harmonic forms on (M, g). As noted in Proposition 2.5.3, the Chern–Simons

1-form i2πTr(Θ) is right-invariant. Also, the form i

2πTr(Θ) lives in the (0, 1) com-

ponent, since THP = Ker Θ. We will not distinguish notationally between the

form i2πTr(Θ) ∈ Ω0,1(P )G and its image in Ω0(M ; g∗P ) under the isomorphism of

complexes in 3.1.7

Ωi,j(P )G π∗←− Ωi(M ; Λjg∗P ).

Also, for any subspace W ⊂ Ω∗(P ), we call W ⊂ Ω∗(P )[|δ|] the image of W under

the inclusion

Ω∗(P ) → Ω∗(P )[|δ|]

(where Ω∗(P ) includes as constant power series.)

Proposition 5.1.2. Given (M,P, g,Θ) with G = U(n) (n ≥ 1), if c1(P ) = 0 ∈

H2(M ; R), then

E1∞,δ =

(R[

i

2πTr(Θ)− δπ∗h]⊕ π∗H1

g(M)

)⊂ Ω1(P )[|δ|],

where h ∈ Ω1(M) is the unique form such that

dh = c1(P,Θ), h ∈ d∗Ω2(M).

Proof. The proof follows in two parts. First, we show that for any element β ∈

H1g(M), π∗β ∈ E1,0

∞,δ. Second, we show that i2πTr(Θ)−δπ∗h ∈ E0,1

∞,δ. As a reminder,

Ei,j∞,δ ⊂ Ωi+j(P )[|δ|] was defined in Definition 4.2.7 by

Ei,j∞,δ

def= ωδ ∈ Ω∗(P )[|δ|] | ω ∈ Ωi,j(P ) ; ωl ∈ E⊥

∞ (l ≥ 1)

dδωδ = d∗δωδ = 0.

π∗H1g(M) = E1,0

∞,δ

76

Page 84: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Let β ∈ H1g(M). Then, π∗β ∈ Ω1,0(P )G is a right-invariant horizontal form on P .

We proceed to show that dδ(π∗β) = d∗δ(π

∗β) = 0. Since π∗β is right-invariant, we

calculate these using the isomorphic complexes (3.3.2) and (3.3.3). For an element

of Ω1,0(P )G = π∗Ω1(M ; R), dδ(π∗•) = π∗(dM•). Since h is closed,

dδ(π∗h) = π∗(dMh) = 0.

Likewise, since h is also coclosed,

d∗δ(π∗h) = π∗(d∗Mh) = 0.

Note that the (2,−1∗) component of d∗δ is 0 here for dimensional reasons. We have

now shown that for any element β ∈ π∗H1g(M), π∗β ∈ Ω1,0(P ) and dδ(π

∗β) =

d∗δ(π∗β) = 0. Since there are no higher order terms (with respect to δ) in the power

series π∗β, we have satisfied all the conditions for π∗β ∈ E1,0∞,δ. Therefore,

π∗H1g(M) ⊂ E1,0

∞,δ.

By the isomorphism of the filtration Ei,jK with the Leray–Serre spectral sequence,

along with the description of E1,0∞ in equation (5.1.1), we see

dimE1,0∞,δ = dimE1,0

∞ = dimH1(M ; R).

Therefore, our inclusion of vector spaces is surjective:

π∗H1g(M) = E1,0

∞,δ.

As a side remark, the relationship π∗Hig(M) = Ei,0

2 holds true in general and is

proved in Proposition 5.2.1. The fact that E1,02 = E1,0

∞ gives us the constant term

in E1,0∞,δ.

77

Page 85: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

R[ i2πTr(Θ)− δπ∗h] = E0,1

∞,δ

First, the existence of an h ∈ Ω1(M) such that dh = c1(P,Θ) is given by the fact

that as a de Rham class,

[c1(P,Θ)] = c1(P ) = 0 ∈ H2(M ; R).

Therefore, c1(P,Θ) ∈ Ω2(M) is an exact form. Using the Hodge decomposition of

Ω1(M), there exists a unique form h such that dh = c1(P,Θ), and h ∈ d∗Ω2(M)

(Corollary 2.4.2).

We proceed by directly computing dδ and d∗δ of i2πTr(Θ) − π∗h. Because this

is a right-invariant form on P , we can calculate using the isomorphic complexes

(3.3.2) and (3.3.3). Under the bi-grading, i2πTr(Θ) ∈ Ω0,1(P )G and π∗h ∈ Ω1,0(P )G.

Furthermore, the derivative of the Chern–Simons 1-form is the first Chern form

(Proposition 2.5.3), so

d(i

2πTr(Θ)− π∗h) = π∗c1(P,Θ)− π∗c1(P,Θ) = 0 ∈ Ω2,0(P )G.

The isometry ρδ is given by multiplying by a power of δ, with that power determined

by the number of horizontal components in a form:

ρδ : Ωi,j(P )[|δ|] −→ Ωi,j(P )[|δ|]

ψ 7−→ δiψ

The rescaled derivative dδ is given by dδ = ρδdρ−1δ . Therefore, if ψ is d-closed, then

ρδψ is dδ-closed. Consequently,

dδρδ

(i

2πTr(Θ)− π∗h

)= dδ

(i

2πTr(Θ)− δπ∗h

)= 0.

Alternatively, one could look at how the components of dδ act in the complex (3.3.2).

The isomorphism Ωi(M ; Λjg∗P ) → Ωi,j(P )G is denoted π∗. We then explicitly

78

Page 86: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

see

(i

2πTr(Θ)− δπ∗h

)=π∗

(dg(

i

2πTr(Θ)) + δd∇(

i

2πTr(Θ)) + δ2

(ιΩ

i

2πTr(Θ)− dMh

))=π∗

(0 + 0 + δ2 (c1(P,Θ)− c1(P,Θ))

)= 0.

While this issue of the bi-grading may seem a bit confusing, one can visualize the

calculation in the picture below. In this, we are thinking of the forms living in the

complex (3.3.2), and the arrows represent the (a priori) non-zero components of

dδ. Finally, we denote the fact that δ2ιΩi

2πTr(Θ) = −δ2dMh by the ± sign in the

image. In other words, these two elements cancel each other. An extra page with

this complex is attached at the end of the thesis. While this may all seem a bit

unnecessary in this calculation, the author finds these pictures extremely helpful in

organizing the calculations in the upcoming Proposition 5.3.1.

0

i2πTr(Θ)

δ2ιΩ

**UUUUUUUUUUUUUUUUU

dg

OO

δ∇ // 0

−δh δdM // ±δ2c1(P,Θ)

We now calculate d∗δ(i

2πTr(Θ)−δπ∗h). Since both forms are right-invariant, when

viewed in the complex (3.3.3) the only non-trivial components of d∗δ(i

2πTr(Θ)−δπ∗h)

are given by

d∗g(i

2πTr(Θ)) and d∗M(h),

or the d0,1∗ and d1,0∗, respectively. Since i∗xi

2πTr(Θ) ∈ g∗P (the restriction of i

2πTr(Θ) ∈

Ω0(M ; g∗P ) to a fiber) is Ad-invariant (Proposition 2.5.3), and Ad-invariant elements

of g∗P are harmonic (Proposition 2.4.4), then i2πTr(Θ) is harmonic when restricted

79

Page 87: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

to the fibers. Therefore,

d∗g(i

2πTr(Θ)) = 0.

By definition, h ∈ d∗Ω2(M), and therefore

d∗Mh ∈ (d∗M)2Ω2(M ; R) = 0.

Hence,

d∗δ(i

2πTr(Θ)− δπ∗h) = π∗(d∗g

i

2πTr(Θ)− δ2d∗Mh) = 0

As above, we can draw this in our complex (3.3.3). This looks like

i2πTr(Θ)

0

0 −δh

δd∗Moo

In particular, the picture makes it easy to identify the potentially non-trivial com-

ponents of d∗δ . It should be noted that we do not have to worry about d2,−1∗h for

dimensional reasons. This does not hold true in the situations for higher-dimensional

forms.

We have now shown that dδ(i

2πTr(Θ) − δπ∗h) = d∗δ(

i2πTr(Θ) − δπ∗h) = 0, and

the constant term i2πTr(Θ) ∈ Ω0,1(P ). In order for i

2πTr(Θ)− δπ∗h to be in E0,1

∞,δ,

all higher order (with respect to δ) terms must be in E⊥∞. In the above discussion,

we showed that

E1,0∞ = π∗H1

g(M).

Our assumption that h ∈ d∗Ω2(M) implies that

h ⊥ H1g(M).

80

Page 88: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This is the only higher-order term, and so we have now verified that

i

2πTr(Θ)− δπ∗h ∈ E1

∞,δ.

By the isomorphism of our filtration on forms Ei,jK with the Leray–Serre spectral

sequence and (5.1.1),

dimE0,1∞ = dimE0,1

∞,δ = 1.

Consequently, the 1-dimensional vector space spanned by i2πTr(Θ)− δπ∗h is equal

to E0,1∞,δ, giving us

R[i

2πTr(Θ)− δπ∗h] = E0,1

∞,δ.

Finally, E1∞,δ = E0,1

∞,δ ⊕ E1,0∞,δ, giving us the desired result

E1∞,δ = R[

i

2πTr(Θ)− π∗h]⊕ π∗H1

g(M).

Theorem 5.1.3. Given (M,P, g,Θ) with G = U(n) (n ≥ 1), if c1(P ) = 0 ∈

H2(M ; R), then for any δ ≥ 0

Ker ∆1gδ

= R[i

2πTr(Θ)− π∗h]⊕ π∗H1

g(M) ⊂ Ω1(P )

where h ∈ Ω1(M) is the unique form such that

dh = c1(P,Θ), h ∈ d∗Ω2(M).

Proof. The proof follows almost directly from Propositions 4.3.5 and 5.1.2. To make

things a bit more explicit, we go ahead and show this directly.

In general, KerLkδ is spanned by Ek

∞,δ under multiplication by elements of R[|δ|]

(Corollary 4.2.8); i.e.

KerLkδ = Ek

∞,δ[|δ|] = Ek∞,δ ⊗R R[|δ|].

81

Page 89: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proposition 5.1.2 says that under our given assumptions,

KerL1δ = E1

∞,δ[|δ|] =

(R[

i

2πTr(Θ)− δπ∗h]⊕ π∗H1

g(M)

)[|δ|].

These power series are actually polynomials, and the formal Lδ coincides with the

analytically defined Lgδon polynomials. Therefore, for any δ ≥ 0,

KerL1gδ

= R[i

2πTr(Θ)− δπ∗h]⊕ π∗H1

g(M).

The isometry

ρδ : (Ωi,j(P ), gδ)→ (Ωi,j(P ), g)

ψ 7→ δiψ

relates the two Laplacians Lgδand ∆gδ

via (4.2.3)

Lgδ= ρδ∆gδ

ρ−1δ .

Also, letting ρδ act formally on power series, we have the formal relationship (ig-

noring the issue of where ρ−1δ is defined)

Lδ = ρδ∆δρ−1δ .

Therefore, ρ−1δ takes elements in KerLgδ

to elements in Ker ∆gδfor any δ > 0. When

defined on a power series or polynomial, then ρ−1δ maps KerLδ to Ker ∆δ. Therefore,

for any δ,

ρ−1δ (

i

2πTr(Θ)− δπ∗h) =

i

2πTr(Θ)− π∗h ∈ Ker ∆gδ

and

ρ−1δ

(δπ∗H1

g(M))

= π∗H1g(M) ⊂ Ker ∆gδ

.

For any δ ≥ 0, KerLgδ∼= Ker ∆gδ

, so the above maps give an inclusion of vector

82

Page 90: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

spaces which is therefore an isomorphism. We conclude that for any δ ≥ 0,

Ker ∆1gδ

=

(R[

i

2πTr(Θ)− π∗h]⊕ π∗H1

g(M)

).

Remark 5.1.4. There are no higher order powers of δ in the basis elements of

Ker ∆gδ. In fact, there is a canonical connection on the bundle Ker ∆1

gδ→ [0, 1]

given by the isomorphism of Ker ∆1gδ

∼= H1(P ; R). In this situation, since the basis

elements of Ker ∆1gδ

have no higher order powers of δ, the bundle Ker ∆1gδ

∼= Ker ∆10×

[0, 1], and the connection is the trivial connection. In other words, the harmonic

representative of a class in H1(P ; R) is independent of the scaling on fibers.

Theorem 5.1.5. Given (M,P, g,Θ) with G = U(n) (n ≥ 1) and c1(P ) 6= 0 ∈

H2(M ; R), then for any δ ≥ 0

Ker ∆1gδ

= π∗H1g(M).

Proof. This follows from the description of E1∞ in (5.1.2), along with the exact same

proof for

π∗H1g(M) ⊂ Ker ∆1

given in the previous Theorem 5.1.3. The fact that c1(P ) 6= 0 implies that we do

not have to describe the cohomology class it generates.

5.2 1- and 2-forms when G is semisimple

If G is a compact semi-simple Lie group, then H1(G; R) = 0, and H2(G; R) = 0

as well. The E2 page of the Leray–Serre spectral sequence is isomorphic to

Ei,j2∼= H i(M ;Hj(G; R)).

83

Page 91: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Consequently, the portion of the E2 page which calculates H1(P ; R) and H2(P ; R)

looks like

0 0 0 0

0 0 0 0

H0(X; R) H1(X; R) H2(X; R) H3(X; R)

(5.2.1)

Since there can be no non-trivial differentials in this portion, we see that the terms

where the sequence collapses in calculating H1(P ; R) and H2(P ; R) are

N = N(1) = N(2) = 2.

Consequently,

E1∞ = E1,0

2∼= H1(M ; R), E2

∞ = E2,02∼= H2(M ; R). (5.2.2)

Before proceeding, we first, we point out a useful and general property.

Proposition 5.2.1. Given (M,P, g,Θ), then for any i,

Ei,02 = π∗Hi

g(M).

Proof. By Definition 4.2.2,

Ei,02 = ψ ∈ Ωi,0(P ) such that ∃ψ1, . . .

dδ(ψ + δψ1 + · · · ) ∈ δ2Ω∗(P )[δ]

d∗δ(ψ + δψ1 + · · · ) ∈ δ2Ω∗(P )[δ]

84

Page 92: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

For β ∈ Hig(M), then π∗β ∈ Ωi,0(P )G, and the only non-zero component of dδ is

d1,0. Under the isomorphism with (3.3.2), we see that because β is closed

dδ(π∗β) = δπ∗(dMβ) = 0.

Drawn in the complex (3.3.2), this looks like

π∗β

0

OO

δdM // 0

Calculating d∗δ(π∗h) using the isomorphic complex (3.3.3), we see that because

β is coclosed,

d∗δ(π∗β) = π∗

(δd∗Mβ + δ2ι∗Ωβ

)= δ2π∗(ι∗Ωβ).

Therefore, d∗δ(π∗β) ∈ δ2Ω∗(P )[δ]. Drawn in the complex (3.3.3), this looks like

0 π∗βδd∗Moo

δ2ι∗Ωhh

We draw the δ2ι∗Ω as dotted since we do not need to worry about it.

We have shown that any element β ∈ Hig(M) satisfies

dδ(π∗β) = 0; d∗δ(π

∗β) ∈ δ2Ω∗(P )[δ].

This gives an inclusion

π∗Hig(M) ⊂ Ei,0

2 . (5.2.3)

85

Page 93: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

By the isomorphism of Ei,jK with the Leray–Serre spectral sequence,

dimEi,02 = dimH i(M ; R).

Therefore, the inclusion (5.2.3) is an equality.

Proposition 5.2.2. Given (M,P, g,Θ) with G semisimple, then

E1∞,δ = π∗H1

g(M) ⊂ Ω1(P )[|δ|].

Proof. The Leray–Serre spectral sequence calculating H1(P ; R) collapses at N =

N(1) = 2, as displayed in (5.2.1). The isomorphism between the Leray–Serre spec-

tral squence and our adiabatic spectral sequence gives

E1∞ = E1,0

2 = π∗H1g(M),

with the second equality given by Proposition 5.2.1. We now show that any element

β ∈ H1g(M) satisfies

dδ(π∗β) = d∗δ(π

∗β) = 0.

This was actually proven in Proposition 5.1.2, and we give the same argument here.

To calculate dδ(π∗β), we see via (3.3.2)

dδ(π∗h) = π∗(δdMh) = 0.

Furthermore, via (3.3.3),

d∗δ(π∗h) = π∗(δd∗Mh) = 0.

There is no δ2ι∗Ωβ term for dimensional reasons. Therefore, we have the stronger

statement that the constant power series π∗β is, in fact, the unique power series

86

Page 94: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

associated to π∗β ∈ E1,0∞ via Theorem 4.2.6, and

E1,0∞,δ = π∗H1

g(M).

Theorem 5.2.3. Given (M,P, g,Θ) with G semisimple, then for any δ

Ker ∆1gδ

= π∗H1g(M) ⊂ Ω1(P ).

Proof. By Proposition 5.2.2 and Corollary 4.2.8, we have

KerL1δ = E1

∞,δ[|δ|] = π∗H1g(M)[|δ|].

Since ρ−1δ is defined on δH1

g(M), Proposition 4.3.5 implies that for any δ ≥ 0,

ρ−1δ

(δπ∗H1

g(M))

= π∗H1g(M) ⊂ Ker ∆1

gδ.

Since this map π∗H1g(M) → Ker ∆1

gδis injective for all δ ≥ 0, and KerL1

∼= Ker ∆1gδ

,

we have an equality of vector spaces

π∗H1g(M) = Ker ∆1

gδ⊂ Ω1(P ).

Proposition 5.2.4. Given (M,P, g,Θ) with G semisimple, then

E2∞ = E2,0

∞ = π∗H2g(M) ⊂ Ω2(P ).

Proof. The Leray–Serre spectral sequence calculating H2(P ; R) collapses at N =

N(2) = 2, as displayed in (5.2.1). This fact, combined with the description of E2

given in Proposition 5.2.1, implies that

E2∞ = E2,0

2 = π∗H2g(M).

87

Page 95: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Theorem 5.2.5. Given (M,P, g,Θ) with G semisimple, then

Ker ∆20 = lim

δ→0Ker ∆2

gδ= π∗H2

g(M) ⊂ Ω2(P ).

Proof. Proposition 5.2.4 says that E2,0∞ = π∗H2

g(M). Corollary 4.3.4 states that the

purely horizontal part of E∞ includes into Ker ∆0. In other words,

E2,0∞ ⊂ Ker ∆2

0.

Since E2∞ = E2,0

∞ ,

π∗H2g(M) = Ker ∆2

0.

5.3 3-forms when G is simple

We now examine the case where G is a compact simple Lie group. We denote

the Chern–Simons 3-form (Definition 2.5.5)

α(Θ) = 〈Θ ∧ Ω〉 − 1

6〈Θ ∧ [Θ ∧Θ]〉 ∈ Ω3(P ).

Under the bi-grading, we write

α(Θ) = α0,3 + α2,1,

where

α0,3 = −1

6〈Θ ∧ [Θ ∧Θ]〉 ∈ Ω0,3(P )

and

α2,1 = 〈Θ ∧ Ω〉 ∈ Ω2,1(P ).

88

Page 96: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This bi-grading follows from the definition of THP = Ker Θ (2.3.2), and the fact that

the curvature Ω ∈ Ω2(M ; gP ) (Proposition 2.3.5). The corresponding 4-dimensional

Chern–Weil form is denoted

〈Ω ∧ Ω〉 ∈ Ω4(M).

The proofs use H1(G; R) = H2(G; R) = 0, and the condition that G is simple

implies H3(G; R) ∼= R. We then only have to worry about the vanishing of a single

characteristic class.

We first describe the portion of the Leray–Serre spectral sequence which calcu-

lates H3(P ; R) (the universal case was discussed in Example 4.1.1). Due to the fact

that H1(G; R) = H2(G; R) = 0, the only non-trivial differential is on the E4 page.

The picture shows E2 = E3 = E4 and the non-trivial d4.

H3(G; R)

d4

))SSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSSS· · ·

0 0 0 0 0

0 0 0 0 0

H0(M ; R) H1(M ; R) H2(M ; R) H3(M ; R) H4(M ; R)

H3(G; R) is isomorphic to R, and the (standard) generator is sent to [〈Ω ∧ Ω〉] ∈

H4(M ; R) by d4 (see Example 4.1.1).

The condition [〈Ω ∧ Ω〉] = 0 ∈ H4(M ; R) is equivalent to the above spectral

89

Page 97: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

sequence collapsing at N = N(3) = 2. This also implies that

E3∞ = E3

2 = E0,32 ⊕ E

3,02∼= H3(G; R)⊕H3(M ; R). (5.3.1)

If [〈Ω∧Ω〉] 6= 0 ∈ H4(M ; R), then we see that d4 is injective on the (0, 3) component,

and hence

E3∞ = E3,0

∞ = E3,02∼= H3(M ; R). (5.3.2)

Proposition 5.3.1. Given (M,P, g,Θ) with G simple and [〈Ω∧Ω〉] = 0 ∈ H4(M ; R),

then

E3∞ = R[α0,3]⊕H3

g(M) ⊂ Ω3(P ).

Furthermore, the unique power series in E3∞,δ given by α0,3 is of the form

α0,3 + δ2α2,1 − δ3π∗h+O(δ3) ∈ Ω3(P )[|δ|],

where h ∈ Ω3(M) is the unique form such that

dh = 〈Ω ∧ Ω〉, h ∈ d∗Ω4(M).

Proof. The order of proof is

• Show E3,0∞ = π∗H3

g(M).

• First attempt at showing exists formal power series of form α0,3 + δ2α1,2 +

O(δ3) ∈ E0,3∞,δ.

• Show there exists formal power series of the form α0,3 + δ2α1,2 +O(δ3) ∈ E0,3∞,δ.

• Show that in the above power series, the (3, 0) component of the δ3 term is h.

E3,0∞ = π∗H3

g(M)

The description of E3,0∞ follows immediately from the fact that the spectral sequence

90

Page 98: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

for H3(P ; R) collapses at N = N(3) = 2, along with the description of E3,02 in

Proposition 5.2.1;

E3,0∞ = E3,0

2 = π∗H3g(M).

Now we move on to the main part of this Proposition. Since N = N(3) = 2, to

show that there exists a formal power series of the form

α0,3 + δ2α2,1 +O(δ3) ∈ E0,3∞,δ,

it suffices to show, by Proposition 4.2.5, that there exist forms h, ψ1,2, such that

(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)∈ δ4Ω4(P )[δ]

d∗δ(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)∈ δ4Ω2(P )[δ]

The fact that both dδ and d∗δ of the polynomial vanish on the order N + 2 implies

uniqueness of the terms of order up to δ2.

First attempt at giving power series

Note that α(Θ) − π∗h ∈ Ω3(P )G (Proposition 2.5.6), which will allow us to calcu-

late in the isomorphic complexes (3.3.2) and (3.3.3). First, we ignore the ψ1,2 and

calculate dδ(α0,3 + δ2α2,1 − δ3π∗h). As a reminder (Proposition 2.5.6),

dα(Θ) = π∗〈Ω ∧ Ω〉 ∈ Ω4,0(P ).

Therefore,

d(α(Θ)− π∗h) = π∗ (〈Ω ∧ Ω〉 − 〈Ω ∧ Ω〉) = 0.

From (4.2.1)

dδ = ρδdρ−1δ = d0,1 + δd1,0 + δ2d2,−1,

91

Page 99: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

and this implies that

dδρδ (α(Θ)− π∗h) = dδ

(α0,3 + δ2α2,1 − δ3π∗h

)= 0. (5.3.3)

Thus, we also know how dδ acts on the two components of α(Θ). In particular, due

to the bi-grading,

d0,1α0,3 = 0, d1,0α0,3 = 0, d1,0α2,1 = 0. (5.3.4)

Under the isomorphism with (3.3.2), we can write this as

(α0,3 + δ2α2,1 − δ3h

)= δ2(ιΩα

0,3 − dgα2,1) + δ4(ιΩα

2,1 − dMh)

= δ4(〈Ω ∧ Ω〉 − 〈Ω ∧ Ω〉) = 0(5.3.5)

Visually, this can be pictured in (3.3.2), where each arrow is a (a priori) non-trivial

component of dδ, and the ± sign is used if two elements cancel.

0

α0,3

OO

//

**UUUUUUUUUUUUUUUUUUU 0

±δ2 12Tr(Ω ∧ [Θ ∧Θ])

δ2α2,1

OO

//

++WWWWWWWWWWWWWWWWWWWWWWWWW 0

−δ3h // ±δ4〈Ω ∧ Ω〉

Now we wish to calculate d∗δ(α0,3 + δ2α2,1 − δ3π∗h). Because we only need that

terms of order δ3 or less are zero, higher powers of δ are ignored. In the picture

below, this is represented by making the corresponding arrows dotted. In particular,

dotted arrows do not indicate the actual image of an element (as a form), only the

bi-grading of where it lives.

92

Page 100: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

When we restrict α0,3 to a fiber (Proposition 2.5.6), it is simply a multiple of the

Ad-invariant form

〈θ ∧ [θ ∧ θ]〉 ∈ Λ3g∗P .

This implies that α0,3 is harmonic when restricted to fibers (Proposition 2.4.4).

Therefore,

d∗gα0,3 = 0.

By assumption, h ∈ d∗MΩ4(M), and therefore d∗Mh = 0. Doing a visual calculation

in the complex (3.3.3) gives ...

α0,3

0

δ3γ1,1 δ2α2,1oo

δ4

ii

0 −δ3hoo

δ5

ii

d∗δ(α0,3 + δ2α2,1 − δ3π∗h) = δ3γ1,1 +O(δ4)

for some element γ1,1 ∈ Ω1(M ; g∗P ). Therefore, we must add in another element to

cancel this term.

Note that since dg is linear over Ωi(M) (3.1.2), then d∗g is linear over Ωi(M) as

well. Therefore, the homology of any vertical column

H•(Ωi(M ; Λ•g∗P ), d∗g) = Ωi(M)⊗H•(Λ

•g∗P , d∗g).

The Lie group G is simple, and therefore H1(Λ•g∗P , d

∗g)∼= H1(Λ•g∗P , dg) = 0. Since

93

Page 101: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

d∗gγ1,1 = 0, this implies that γ1,1 ∈ d∗gΩ1(M ; Λ2g∗P ). Define ψ1,2 ∈ Ω1(M ; Λ2g∗P ) to

be the unique element, under the Hodge decomposition of Λ2g∗P , such that

d∗gψ1,2 = γ1,1; dgψ

1,2 = 0.

Successfully showing existence of power series

Now we show that

(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)∈ δ4Ω4(P )[δ]

d∗δ(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)∈ δ4Ω2(P )[δ]

The dδ calculation follows in the complex (3.3.2) as before, but with the added

information that dgψ2,1 = 0:

(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)= δ2(ιΩα

0,3 − dgα2,1)− δ3(dgψ

1,2) +O(δ4)

= 0 +O(δ4).

0

α0,3

OO

//

++WWWWWWWWWWWWWWWWWWWWWWWW 0

−δ3ψ1,2

OO

δ4//

δ5

,,

±δ2 12Tr(Ω ∧ [Θ ∧Θ])

δ2α2,1

OO

//

++WWWWWWWWWWWWWWWWWWWWWWWWW 0

−δ3h // ±δ4〈Ω ∧ Ω〉

94

Page 102: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Likewise the d∗δ follows as before, but with the d∗gψ1,2 cancelling the d∗∇α

2,1 term:

d∗δ(α0,3 + δ2α2,1 − δ3(π∗h+ ψ1,2)

)= d∗gα

0,3 − δ3d∗Mh+ δ3(d∗∇α2,1 − d∗gψ1,2) +O(δ4)

= 0 +O(δ4).

α0,3

0 −δ3ψ1,2

δ4oo

±δ3γ1,1 δ2α2,1oo

δ4

jj

0 −δ3hoo

δ5jj

To show that a power series of the form α0,3 + δ2α2,1 + O(δ3) ∈ E0,3∞,δ, we must

also show that α2,1 ∈ E⊥∞. In (5.3.1) we showed that

E3∞ = E0,3

2 ⊕ E3,02 .

Therefore, α2,1 ∈ Ω2,1(P ) ⊂ E⊥∞. By Proposition 4.2.5, there exists a formal power

series in E3∞,δ whose terms of order δ ≤ 2 are equal to the above power series. We

have then proved the existence of a power series of the form

α0,3 + δ2α2,1 +O(δ3) ∈ E0,3∞,δ.

This also implies that the constant term α0,3 ∈ E0,3∞ . By (5.3.1),

dimE0,3∞,δ = dimH3(G; R) = 1,

95

Page 103: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

and thus

E0,3∞ = R[α0,3].

Properties of h term

We have proved there exist forms φ3, φ4, . . . ∈ Ω3(P ) such that

dδ(α0,3 + δ2α2,1 + δ3φ3 + δ4φ4 + · · · ) = 0 (5.3.6)

d∗δ(α0,3 + δ2α2,1 + δ3φ3 + δ4φ4 + · · · ) = 0 (5.3.7)

We now investigate the (3, 0) component of φ3, which we denote φ3,0. We first wish

to see that φ3,03 is right-invariant. Analyzing the δ3 component of (5.3.6), we see

that

δ3(d1,0α2,1 + d0,1φ3,03 ) = 0.

From earlier calculations in this proof (5.3.4), d1,0α2,1 = 0. Therefore,

d0,1φ3,03 = 0,

which implies that φ3,03 is constant along the fibers, and hence (Lemma 3.1.3)

φ3,03 ∈ Ω3(P )G.

Since φ3,03 is right-invariant, we can use the simpler complex (3.3.2) to see dM and

d∗M . By analyzing the (4, 0) component of δ4 in the first equation (5.3.6),

δ4(ιΩα2,1 + dM(φ3,0

3 )) = 0.

Therefore (by (5.3.5)), dM(φ3,03 ) = −〈Ω ∧ Ω〉. By analyzing the (2, 0) component of

δ4 in the second equation (5.3.7),

δ4d∗M(φ3,03 ) = 0.

96

Page 104: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Furthermore, we require that φ3 ⊥ E∞, so that we have the unique element in E3∞,δ

according to Theorem 4.2.6. Since E3,0∞ = π∗H3

g(M), then

φ3,03 ⊥ H3

g(M).

This uniquely characterizes the form φ3,03 , which we denote −π∗h. Therefore, h ∈

Ω3(M) with

dh = −〈Ω ∧ Ω〉, h ∈ d∗Ω4(M).

The power series is of the form

α0,3 + δ2α2,1 − δ3π∗h+O(δ3) ∈ E0,3∞,δ, (5.3.8)

where the remainder of the δ3 term is perpendicular to Ω3,0(P ).

Theorem 5.3.2. Given (M,P, g,Θ) with G simple and [〈Ω∧Ω〉] = 0 ∈ H4(M ; R),

then

Ker ∆30 = lim

δ→0Ker ∆3

gδ= R[α(Θ)− π∗h]⊕ π∗H3

g(M) ⊂ Ω3(P ).

Here, α(Θ) is the Chern–Simons 3-form, and h ∈ Ω3(M) is the unique form such

that

dh = 〈Ω ∧ Ω〉, h ∈ d∗Ω4(M).

Proof. By Proposition 5.3.1, E0,3∞,δ is generated by an element of the form

α0,3 + δ2α2,1 − δ3π∗h+O(δ3),

where the remainder of the δ3 term is orthogonal to Ω3,0(P ). This implies that ρ−1δ

is defined on the above power series, since ρ−1δ divides by δi where i is the horizontal

part of the bi-grading. By Proposition 4.3.3,

(ρ−1

δ (α0,3 + δ2α2,1 − δ3π∗h+O(δ3)))

δ=0∈ Ker ∆3

0,

97

Page 105: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

and

(ρ−1

δ (α0,3 + δ2α2,1 − δ3π∗h+O(δ3)))

δ=0=

(α0,3 + α2,1 − π∗h+O(δ)

)δ=0

= α(Θ)− π∗h.

Proposition 5.3.1 also states that E3,0∞ = π∗H3

g(M). Using Corollary 4.3.4, this gives

us

π∗H3g(M) ⊂ Ker ∆3

0.

We have now given an injective map E3∞ → Ker ∆3

0, and since these spaces are

isomorphic, we have the equality

Ker ∆30 = R[α(Θ)− π∗h]⊕ π∗H3

g(M).

Remark 5.3.3. In the analogous situation for a U(n)-bundle, we saw that the space

of harmonic 1-forms on P is independent of scaling factor δ. In the above proof for

harmonic 3-forms, one sees the possibility of higher order δ terms in Ker ∆δ. In

fact, these terms do appear in explicit calculations, and the description of harmonic

forms in terms of the Chern–Simons form only holds in the adiabatic limit. It is

not known what information the higher order δ terms contain.

Theorem 5.3.4. Given (M,P, g,Θ) with G simple and [〈Ω∧Ω〉] 6= 0 ∈ H4(M ; R),

then

Ker ∆30 = lim

δ→0Ker ∆3

gδ= π∗H3

g(M) ⊂ Ω3(P ).

Proof. This follows from (5.3.2) and the first part of Proposition 5.3.1:

E3∞ = E3,0

2 = π∗H3g(M).

98

Page 106: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Combining this with Corollary 4.3.4 gives us the inclusion

π∗H3g(M) ⊂ Ker ∆3

0,

which is an equality since both vector spaces have the same dimension.

99

Page 107: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 6

CANONICAL FORMS ASSOCIATED TO LIFTS OF THE STRUCTURE

GROUP

In this section, we define a string structure in an analogous way to a spin

structure: it is a particular type of cohomology class on the total space P . A

string structure is a class S ∈ H3(P ; Z) that restricts to the standard generator of

H3(Spin(n); Z) on the fibers (n ≥ 5). If we introduce a connection on P and a

metric on M , we can take the harmonic representative of the (R-image) of S in the

adiabatic limit. This harmonic representative is the Chern–Simons 3-form minus

a 3-form H ∈ Ω3(M). This form H has properties similar to the Chern–Simons

3-form. We should note that in this construction, we only consider the R-image of a

string structure and lose all information about the torsion in H3(P ; Z). It is hoped

that this construction can be improved so that the torsion information is not lost.

Such a construction would most likely use some form of differential cohomology.

6.1 Generalities about G′ structures

So far, we have dealt with principal G-bundles over a space M , where G is a

compact Lie group. However, the notion of a principal bundle makes sense for more

general topological groups. The general theory of fiber bundles tells us that for a

topological group G, there is a universal G-bundle EG → BG, where BG is called

the classifying space of G. In other words, any G-bundle over M is the pullback

100

Page 108: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

of a classifying map M → BG. The spaces EG and BG are characterized by the

following: given a contractible space EG on which G acts freely, let BGdef= EG/G.

The bundle G → EG → BG is then the universal G-bundle, and is unique up to

homotopy. So, any G-bundle P → M is the pullback of a classifying map in the

following diagram:

G

G

P

f∗ // EG

X

f // BG

This gives us a bijection

G bundles over M ←→Map(M,BG).

In fact, any two bundles over M are isomorphic if and only if their classifying maps

are homotopic (f ' f ′). This gives a bijection

G bundles over M/iso←→ [M,BG],

where the brackets denote homotopy classes of maps.

Given such a G-bundle, one often wants to make the bundle compatible with

some additional structure. Common examples include a choice of orientation, spin

structure, or complex structure. These can often be phrased in terms of reducing or

“lifting” the structure group. Given a group homomorphism G′ → G, this induces a

map BG′ → BG. Given a principal G-bundle P , then a “lift of the structure group

to G” is a choice of lift f

BG′

M

f //

ef <<yy

yy

BG

101

Page 109: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

The lift f induces a G′ bundle which “covers” P in the following sense. Given

G′

xxxx

xxxx

x

G

G

EG′

xxxxxxxx

P

f∗ // EG

BG′

xxxxxxxx

Mf //

ef 22s

pn

l j h

BG

we obtain

G′ //

G

P ′ //

BBB

BBBB

B P

M

The most basic examples are given by inclusions G′ → G. In this case, such a lift

is often called a reduction of the structure group to G′. For example, An orientation

is a reduction of the structure group to SO(n) → O(n). Another fundamental but

more subtle example is a spin structure. On a principal SO(n)-bundle P → M , a

choice of a spin structure is a lift

BSpin(n)

M

f //

ef ::uu

uu

uBSO(n)

Here, the map BSpin(n) → BSO(n) is induced from the double cover Spin(n) →

SO(n).

For simplicity, we restrict to n ≥ 3, which is the stable range for π1(SO(n)).

102

Page 110: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Then, the homomorphism Spin(n)→ SO(n) is special in that

πi(Spin(n)) =

πi(SO(n)) i > 1

0 i = 0, 1

In other words, Spin(n) is the 1-connected cover of SO(n). Furthermore, in the lan-

guage of bundles, we know that Spin(n) is a π1(SO(n)) ∼= Z/2 bundle over SO(n).

Therefore, up to homotopy, Spin(n) is simply the pullback of a map SO(n)→ BZ/2.

Z/2

Z/2

Spin(n)

f∗ // EZ/2

SO(n)

f // BZ/2

Since Z/2 is a discrete group, as a space it is the Eilenberg-MacLane spaceK(Z/2, 0).

Eilenberg-MacLane spaces are characterized by the property

πi(K(H,n)) =

H i = n

0 i 6= n

for any abelian group H and n ∈ N. They are also the representing spaces for the

ordinary cohomology spectrum, giving

Hn(X;H) = [X,K(H,n)].

They are also topological groups and are compatible with the B functor in the sense

that

BK(H,n) ' K(H,n+ 1).

Viewed in this language, we see that up to homotopy, Spin(n) is the pullback of a

103

Page 111: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

map SO(n)f→ BK(Z/2, 0) ' K(Z/2, 1)

K(Z/2, 0)

K(Z/2, 0)

Spin(n)

f∗ // EK(Z/2, 0)

SO(n)

f // K(Z/2, 1)

The map SO(n)f→ K(Z/2, 1) gives the generator of H1(SO(n); Z/2).

Applying B to the above bundle gives a description of BSpin(n) as the pullback

of Bf in the diagram

K(Z/2, 1)

K(Z/2, 1)

BSpin(n)

Bf∗ // EK(Z/2, 1)

BSO(n)

Bf // K(Z/2, 2)

The mapBf represents the universal second Stieffel-Whitney class w2 ∈ H2(BSO(n); Z/2).

One may feel this abstraction is unnecessary, as we have explicit descriptions of

Spin(n) without all this topology language. However, it is not clear how to best

think about the 3-connected cover of Spin(n), which we denote String(n) (when

n ≥ 5). Yet, there is interesting cohomological information that is seen by only

considering this topological group up to homotopy.

We now describe a standard generalization of the above construction and think

about lifts of the structure group to a topological group G′, where G′ is, up to

homotopy, the homotopy fiber of a map G→ K(H,n). Note that we always assume

the existence of the topological groups G and G′, and then consider their homotopy

type.

Let K be a path-connected CW complex, and fix a basepoint k0 ∈ K. Then,

104

Page 112: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

define PK, the path-space of K, along with a map π : PK → K by

PKdef= γ : [0, 1]→ K|γ(0) = k0

π−→ K

γ 7−→ γ(1)

While the pathspace PK is contractible, the fiber of π : PK → K at a point k1 ∈ K

is

PKk1 = π−1(k1) = γ : [0, 1]→ K|γ(0) = k0, γ(1) = k1,

which is homotopy equivalent to the pointed loop space

ΩKdef= γ : [0, 1]→ K|γ(0) = γ(1) = k0.

Definition 6.1.1. Given a map f : Y → K, we define the homotopy fiber Y ′ to

be the pullback of f in the diagram

ΩK

ΩK

Y ′

f∗ // PK

Y

f // K

Proposition 6.1.2. Suppose Y ′ is the homotopy fiber of Y → K(H,n). Then,

πi(Y′) ∼= πi(Y ); i 6= n, n− 1

Furthermore, if Y → K(H,n) is the Hurewicz image of πn(Y ), then

πi(Y′) =

πi(Y ) i 6= n

0 i = n

Proof. The proof follows directly from the long exact sequence of homotopy groups

associated to the fibration.

105

Page 113: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proposition 6.1.3. There is a natural isomorphism πk+1BG→ πkG induced from

the fibration G→ EG→ BG.

We now consider the situation where G′ is, up to homotpy, the homotopy fiber

of a map G→ K(H,n), where K(H,n) is an Eilenberg-MacLane space and H some

abelian group. Let BG′ be the corresponding classifying space, which is the homo-

topy fiber of the induced map BG→ K(H,n+ 1). In other words, BG′ is obtained

from “killing” a cohomology class. Let C ∈ Hn(G;H) and B[C] ∈ Hn+1(BG;H)

denote the cohomology classes associated to the maps C : G → K(H,n) and

BC : BG→ K(H,n+ 1).

Proposition 6.1.4. Let P → M be a principal G-bundle, classified by a map f :

M → BG. The following are equivalent:

• a homotopy class of lifts f : M → BG′ of the fixed classifying map f .

• a cohomology class G′ ∈ Hn(P ;H) such that i∗G′ = C ∈ Hn(G;H).

Proof. The proof follows from the fact that iterating the homotopy fiber construction

gives us the sequence

· · · → G′ → G→ K(H,n)→ BG′ → BG→ K(H,n+ 1)

where any two consecutive maps are a fibration. The homotopy fiber of BG′ → BG

is K(H,n). The existence of f allows us to perform a lift of maps between fibrations

G

zzttttttttttt

G

//

22

qo

mk i g f

G

K(H,n)

zzttttttttt

P

f∗ //

ef∗ 33q

om

k i g

EG

BG′

yysssssssss

Mf //

ef 22q

om

k i g e

BG

106

Page 114: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Therefore, we see a map f ∗ : P → K(H,n) that restricts to our preferred map

G → K(H,n) on the fibers. This gives an element G′ ∈ Hn(P ;H) that restricts to

the canonical element C on the fibers. Any two lifts f which are homotopic induce

the same cohomology class.

Conversely, a cohomology class G′ ∈ Hn(P ;H) restricting to C on the fibers

gives, up to homotopy, the map f ∗ : P → K(H,n) restricting to G → K(H,n) on

the fibers. Such a map f ∗ induces f .

Proposition 6.1.5. Assume H i(G;H) = 0 for i < n. Given a principal bundle

G→ P →M as in Proposition 6.1.4. Then

1. a G′ structure exists if and only if BCf ' ∗, or equivalently, the characteristic

class of P corresponding to BC is 0.

2. the space of G′ structures is a torsor for Hn(M ;H).

Proof. The proof follows directly from the Leray–Serre spectral sequence for the

fibration. The E2 = · · · = En+1 pages are equal, and the only non-trivial differential

is the following dn+1

Hn(G;H)

dn+1

''PPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPPP

0 0 · · ·

......

0 0 · · ·

H0(M ;H) H1(M ;H) · · · · · · Hn(M ;H) Hn+1(M ;H)

107

Page 115: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

This induces the exact sequence

0 // Hn(M ;H) π∗ // Hn(P ;H) i∗ // Hn(G;H)dn+1// Hn+1(M ;H)

G′ ? // C

// BC(P )

Therefore, there exists such aG′ structure G′ if and only if C maps to 0. Furthermore,

any two such G′ structures will differ by an element of Hn(M ;H).

Proposition 6.1.6. Assume G is (n− 1) connected (i.e. that πi(G) = 0 for i < n)

and that C is the Hurewicz image of πnG. Then, we have the additional equivalence

for P →M classified by f : M → BG:

• a homotopy class of lifts f : M → BG′ of the classifying map f .

• a trivialization (global frame) of P on the n skeleton of M that extends to the

n+ 1 skeleton. Trivializations are considered up to homotopy.

Proof. The hypotheses imply that G′ is the n-connected cover of G, and BG′ is

the (n + 1)-connected cover of BG. Therefore, given a lift f : X → BG′, f is

nullhomotopic on the (n + 1)-skeleton of X, and any two nullhomotopies on the

n-skeleton are homotopic to each other. Projecting these homotopies to X → BG,

we see that the first condition implies the second. Conversely, if classifying map

f : X → BG is nullhomotopic on the (n + 1)-skeleton, then the map C f is

nullhomotopic on all of X. A particular choice of nullhomotopy of C f gives a lift

to X → PK(H,n+ 1), which in turn pulls back to a lift f : X → BG′.

K(H,n)

K(H,n)

BG′

C∗ // PK(H,n+ 1)

X

f //

ef ;; 44

BGC // K(H,n+ 1)

108

Page 116: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

6.2 U structures on U(n)-bundles

Suppose we have a U(n)-principal bundle P over a manifold M . For n ≥ 1,

π1(U(n)) ∼= H1(U(n); Z) ∼= Z,

and we consider the universal cover U(n)→ U(n). Up to homotopy, this is just the

fiber of the standard map U(n)→ K(Z, 1) that generates the first cohomology:

U(n)→ U(n)→ K(Z, 1).

Remark 6.2.1. The inclusion U(1) → U(n) gives an isomorphism on the first

cohomology, and the standard orientation of U(1) picks out a canonical generator

of H1(U(1); Z). This gives the standard generator for H1(U(n); Z).

Remark 6.2.2. The group U(n) has an explicit description as the semi-direct prod-

uct SU(n) n R.

Definition 6.2.3. (n ≥ 1) A U structure on a U(n)-bundle P →M is a cohomology

class U ∈ H1(P ; Z) such that i∗U ∈ H1(U(n); Z) is the standard generator, where

i : U(n) → P is the fiber-wise inclusion.

Proposition 6.2.4. A U structure is equivalent to

• a homotopy class of lifts f BU(n)

M

ef <<yy

yy

y f // BU(n)

• a homotopy class of trivializations of P on the 1-skeleton that extend to the

2-skeleton of M .

109

Page 117: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

A U structure exists if and only if c1(P ) = 0 ∈ H2(M ; Z). Furthermore, the space

of U structures is a torsor for H1(M ; Z).

Proof. These statements follow from Propositions 6.1.4, 6.1.5, and 6.1.6.

Theorem 6.2.5. Let P be a principal U(n)-bundle with connection Θ over a Rie-

mannian manifold (M, g). If U ∈ H1(P ; Z) is a choice of U-structure, then for any

δ ≥ 0,

[U]gδ=

i

2πTr(Θ)− π∗Hg,Θ,eU ∈ Ω1(P ).

The form Hg,Θ,eU ∈ Ω1(M) satisfies dHg,Θ,eU = c1(P,Θ) and d∗Hg,Θ,eU = 0. Further-

more, if the U structure is changed by ξ ∈ H1(M ; Z), then

[U + π∗ξ]gδ= [U]gδ

+ π∗[ξ]g.

Proof. The existence of a U structure implies that c1(P ) = 0 ∈ H2(M ; R). There-

fore, Theorem 5.1.3 says that

Ker ∆1gδ

= R[i

2πTr(Θ)− π∗h]⊕ π∗H1

g(M),

where dh = c1(P,Θ) and h ∈ d∗Ω2(M). The cohomology class U restricts to the

generator in H1(U(n); Z). Therefore, the harmonic representative [U]gδmust be of

the form

[U]gδ= 1[

i

2πTr(Θ)− π∗h] + π∗(h′)

where h′ ∈ H1g(M). Setting Hg,Θ,eU = h− h′, we see that

[U]gδ=

i

2πTr(Θ)− π∗h+ π∗h′ =

i

2πTr(Θ)− π∗Hg,Θ,eU.

Finally, any two U structures U1, U2 are related by

U1 − U2 = π∗ξ

110

Page 118: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

for some ξ ∈ H1(M ; Z). By the description of harmonic forms on P , we see that

the harmonic representative of π∗ξ ∈ H1(P ; R) is the pullback of [ξ]g ∈ H1g(M).

6.3 Spinc structures on SO(n)-bundles

For the following, assume n ≥ 3. Note that

H2(SO(n); Z) ∼= H1(SO(n); Z) ∼= Z/2.

The group Spinc(n) is, up to homotopy, the homotopy fiber of the map SO(n) →

K(Z, 2) generated by the non-trivial cohomology class. Note that spinc is not simply-

connected, but instead is an S1 extension of SO(n) (since K(Z, 2) ' BU(1)).

Definition 6.3.1. (n ≥ 3). A spinc structure on a SO(n)-bundle P → M is a

cohomology class c ∈ H2(P ; Z) such that i∗c 6= 0 ∈ H2(SO(n); Z) ∼= Z/2, where

i : SO(n) → P is the fiber-wise inclusion.

Proposition 6.3.2. A spinc structure on P →M is equivalent to a homotopy class

of lifts f

BSpinc(n)

M

ef ::tt

tt

t f // SO(n)

A spinc structure exists if and only if W3(P ) = 0 ∈ H3(M ; Z), where W3 is the third

integral Stieffel-Whitney class (obtained by the Bockstein of w2). The space of spinc

structures is a torsor for H2(M ; Z).

Proof. These statements follow from Propositions 6.1.4 and 6.1.5, along with the

fact that in the universal case, the non-trivial element in H2(SO(n); Z) transgresses

to W3 ∈ H3(BSO(n); Z) via the Leray–Serre spectral sequence.

111

Page 119: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Theorem 6.3.3. (n ≥ 3) Let P be a principal SO(n)-bundle with connection Θ

over a Riemannian manifold (M, g). If c ∈ H2(P ; Z) is a choice of spinc structure,

then

[c]0 = limδ→0

[c]gδ= π∗Hg,Θ,c ∈ Ω2(P ).

Here, Hg,Θ,c ∈ H2g(M). Furthermore, if the spinc structure is changed by ξ ∈

H2(M ; Z), then

[c+ π∗ξ]0 = [c]0 + π∗[ξ]g.

Proof. Since H2(SO(n); R) = 0, Theorem 5.2.5 says that

Ker ∆20 = π∗H2

g(M)

and hence [c]0 ∈ π∗H2g(M). Furthermore, for ξ ∈ H2(M ; Z), the adiabatic harmonic

representative of the class pulled back to P is merely the pullback of the harmonic

representative on M :

[π∗ξ]0 = π∗[ξ]g ∈ Ω2(P ).

The above Theorem is interesting in the following way. A spinc structure, topo-

logically, is a choice of a particular complex line bundle L→ P , characterized by its

first Chern class c ∈ H2(P ; Z). To any spinc structure, there is an associated line

bundle λ on M . This is determined by requiring that

π∗c1(λ) = 2c ∈ H2(P ; Z).

In usual geometric applications, one must also choose a connection on the associated

line bundle λ. However, if we introduce a metric g on M (along with a connection

on P , which is usually already determined by g), then taking the harmonic rep-

resentative of c in the adiabatic limit produces a canonical 2-form Hg,c ∈ Ω2(M).

112

Page 120: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Furthermore, we see that π∗c1(λ) = 2[H] ∈ H2(P ; R), and hence 2Hg,c ∈ Ω2(M) is

the curvature for some connection on λ. Furthermore, Hg,c is harmonic, and so picks

out a class of connections on λ with energy-minimizing curvature. In applications,

it is often preferable to use such connections.

6.4 String structures on Spin(n)-bundles

For now, assume n ≥ 5. We will treat the case n ≤ 4 in a moment. For n ≥ 5,

Spin(n) is 2-connected and

π3(Spin(n)) ∼= H3(Spin(n); Z) ∼= Z.

Definition 6.4.1. String(n) is a topological group that is the 3-connected cover of

Spin(n). Up to homotopy, it is the homotopy fiber of the canonical map Spin(n)→

K(Z, 3).

The group String(n) does exist, and one construction can be found in section

5.4 of [ST]. It should be noted that String(n) cannot be a compact Lie group, so

any known model forces us to deal with more sophisticated ideas. Here, we can

avoid this issue by simply assuming its existence and then thinking about it only up

to homotopy. In particular, whatever model of String(n) is used, a string structure

will still give you the following cohomology class.

Definition 6.4.2. (n ≥ 5) A string structure on a Spin(n)-bundle P → M is

a cohomology class S ∈ H3(P ; Z) such that i∗S ∈ H3(Spin(n); Z) is the standard

generator, where i : Spin(n) → P is the fiber-wise inclusion.

Proposition 6.4.3. A string structure on P →M is equivalent to

1. a homotopy class of lifts f BString(n)

M

ef 99ssssss f // BSpin(n)

113

Page 121: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

2. a homotopy class of trivializations of P on the 3-skeleton of M which extend

to the 4-skeleton.

Furthermore, a string structure exists if and only if p1

2(P ) = 0 ∈ H4(M ; Z), and the

space of string structures is a torsor for H3(M ; Z).

Proof. These statements follow from Propositions 6.1.4, 6.1.5, and 6.1.6, along with

the fact that in the universal case, the generator of H3(Spin(n); Z) transgresses to

p1

2∈ H4(BSpin(n); Z) in the Leray–Serre spectral sequence.

The notion of a string structure is based on the obstruction p1

2(P ) ∈ H4(M ; Z).

As discussed at the end of Chapter 2, the p1

2class is a stable characteristic class.

The definition of String(n) holds for all n, but it is only for n ≥ 5 that we have

the nice description of String(n) as being the 3-connected cover of Spin(n). This is

due to the fact that π3(Spin(n)) ∼= H3(Spin(n); Z) stabilizes at n = 5. This same

thing happens when defining the groups Spin(n) and spin structures. The groups

Spin(n) are always double-covers of SO(n), but only for n ≥ 3 does this have the

description of being the universal cover.

Therefore, we say that a string structure for n ≤ 4 is a cohomology class S ∈

H3(P ; Z) such that i∗xS ∈ H3(Spin(n); Z) is the image of the standard generator

of H3(Spin(5); Z) under the inclusion Spin(n) → Spin(5). Alternatively, given

a Spin(n)-bundle with n ≤ 4, we can look at the induced Spin(5)-bundle and

talk about string structures on it. Since this is also how the groups String(n) are

characterized for n ≤ 4, a string structure is still equivalent to a homotopy class of

lifts of the classifying map to BString(n), and it is also equivalent to a homotopy

class of trivializations of the stable bundle of P on the 3-skeleton which extends to

the 4-skeleton.

114

Page 122: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Theorem 6.4.4. Let P be a principal Spin(n)-bundle with connection Θ over a

Riemannian manifold (M, g). If S ∈ H3(P ; Z) is a choice of string structure, then

the adiabatic limit of the harmonic representative is given by

[S]0 = limδ→0

[S]gδ= α(Θ)− π∗Hg,Θ,S ∈ Ω3(P ).

The form Hg,Θ,S ∈ Ω3(M) has the properties dHg,Θ,S = p1

2(P,Θ), d∗Hg,Θ,S = 0.

Furthermore, if the string structure is changed by ξ ∈ H3(M ; Z), then

[S + π∗ξ]0 = [S]0 + π∗[ξ]g.

Proof. The existence of a string structure implies that p1

2(P ) = 0 ∈ H4(M ; R).

Therefore, Theorem 5.3.2 says that

Ker ∆30 = R[α(Θ)− π∗h]⊕ π∗H3

g(M),

where dh = p1

2(P,Θ) and h ∈ d∗Ω4(M). The cohomology class S restricts to the

generator in H3(Spin(n); Z). Therefore, the harmonic representative [S]0 must be

of the form

[S]0 = 1[α(Θ)− π∗h] + π∗(h′)

where h′ ∈ H3g(M). Setting H = h− h′, we see that

[S]0 = α(Θ)− π∗h+ π∗h′ = α(Θ)− π∗Hg,Θ,S.

Finally, for ξ ∈ H4(M ; Z), the adiabatic harmonic representative of the class pulled

back to P is merely the pullback of the harmonic representative on M :

[π∗ξ]0 = π∗[ξ]g ∈ Ω4(P ).

It may be tempting to think, at least rationally, that there is a canonical string

115

Page 123: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

structure determined by the real cohomology class [α(Θ) − π∗h]. However, this

is not necessarily an integral class. In general, some closed form from the base

must be added so that the resulting form has integral periods. Consequently, the

canonical form Hg,Θ,S ∈ Ω3(M) associated to a string structure normally does not

have integral periods. In fact, this has a natural interpretation in terms of classical

Chern–Simons theory.

A global section p : Mπ→ P gives an isomorphism of principal bundles

P ∼=pM × Spin(n).

Via the Kunneth formula, along with the fact thatH1(Spin(n); Z) = H2(Spin(n); Z) =

0, the above isomorphism induces the following isomorphism on integral cohomology,

H3(P ; Z) ∼= H3(M ; Z)⊕H3(Spin(n); Z)

S↔ (0, 1Spin)

and we say the string structure determined by p is the element S ∈ H3(P ; Z) which

corresponds to the pullback of the generator 1Spin ∈ H3(G; Z). In the following

proposition, we let HS = Hg,Θ,S.

Proposition 6.4.5. Let HS be given by (M,P, g,Θ, S). If p : M → P is a global

section with induced string structure S ∈ H3(P ; Z), then

[p∗α(Θ)−HS] = 0 ∈ H3(M ; R).

Consequently, if d∗(p∗α(Θ)) = 0, then p∗α(Θ) = Hξ.

Proof. First,

d(p∗α(Θ)−HS) =p1

2(Θ)− p1

2(Θ) = 0.

116

Page 124: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Therefore, [p∗α(Θ)−HS] ∈ H3(M ; R). We now show that it evaluates as 0 on any

3-cycle X in M . Let X ∈ Map(X3,M) where X3 is a 3-dimensional manifold.

We use the notation∫

Xto denote evaluation on the 3-cycle X. Then, under the

isomorphism P ∼= M ×G induced by the global section p,

p(X) ⊂M × pt ⊂ P.

Since ξ ∈ H3(P ; Z) is zero on 3-cycles in M ⊂ P ,

〈ξ, p(X)〉 = 0.

This induces the relationship in real cohomology

0 =

∫p(X)

[ξ]0 =

∫p(X)

(α(Θ)− π∗HS) =

∫X

(p∗α(Θ)−HS) .

This holds for an arbitrary 3-cycle, and hence [p∗α(Θ) − HS] = 0 ∈ H3(M ; R).

Furthermore,

d∗HS = 0

by Theorem 6.4.4. Therefore, if d∗(p∗α(Θ)) = 0, then p∗α(Θ)−HS is harmonic and

consequently equals 0. In other words, under the Hodge decomposition, the forms

p∗α(Θ) andHS are equal on the coexact and harmonic components. If d∗(α(Θ)) = 0,

then the forms are equal.

6.5 Canonical forms on base from G′ structures

Let P → M be a principal G-bundle, where G is appropriate to our situation.

If Met(M) is the space of Riemannian metrics on M , and AP is the affine space of

connections on P , then we have canonical maps given by looking at the term from

117

Page 125: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

the base in the previous adiabatic limit constructions.

U Structures ×Met(M)×AP −→ Ω1(M)

Spinc Structures ×Met(M)×AP −→ Ω2(M)

String Structures ×Met(M)×AP −→ Ω3(M)

Also, these maps are equivariant with respect to the action of H1(M ; Z), H2(M ; Z),

and H3(M ; Z) respectively, where H i(M ; Z) acts freely and transitively on the space

of Structures by π∗, and acts on Ωi(M) by the addition of harmonic representatives.

Often the bundle P is determined by the metric g. In particular, frame bundles

such as Spin(M) depend on the metric and have a canonical connection determined

by the metric g. In the case of a manifold M with spin structure, we obtain a map

String Structures ×Met(M) −→ Ω3(M)

by applying the construction to (M,Spin(M), g,ΘLC), where ΘLC is the Levi-Civita

connection.

The above Proposition 6.4.5 is quite helpful if one wants to calculate Hg,S, and

the bundle is trivializable. Instead of having to find a harmonic representative of S,

which is a difficult global problem (even in the adiabatic limit), one can calculate the

Chern–Simons form on a “good” choice of global frame. Given a frame p : M → P

(or local frames), the connection can be pulled back to a Lie-algebra valued 1-form

on the base. This is how local calculations in geometry are usually performed. We

will denote this Θp = p∗Θ and Ωp = p∗Ω. Then,

p∗α(Θ) = 〈Θp ∧ Ωp〉so(g) −1

6〈Θp ∧ [Θp ∧Θp]〉so(g).

In the following calculation, G will be the base space, and the fiber will be Spin(g) ∼=

Spin(n) where n = dimG.

118

Page 126: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proposition 6.5.1. Let G be any compact Lie group with bi-invariant metric g.

Let Spin(G) → G be a principal Spin(g) frame bundle for TG with the Levi-

Civita connection Θg. If L ∈ H3(Spin(G); Z) is the string structure induced by

left-invariant framing L, and R ∈ H3(Spin(G); Z) is the string structure induced by

right-invariant framing R, then

HL = − 1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g)

HR =1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g)

where θ : TG→ g is the Maurer-Cartan 1-form.

Proof. Let g denote the Lie algebra of left-invariant vector fields on G. Under the

left-invariant framing, we have the isomorphism

TG ∼= G× TeG,

where e ∈ G is some point. Any constant map G→ TeG in the above isomorphism

is a left-invariant vector field, so it is convenient to compose with θ. This gives the

trivialization

TG≈−→ G× TeG

G×θ−→ G× g.

We now calculate the connection and curvature forms on this frame, denoted ΘL

and ΩL. Let X, Y, Z be left-invariant vector fields, and hence constant sections of

G× TeG. Then, the Levi-Civita connection ∇ on TG is given by the formula

2〈∇XY, Z〉 =〈[X, Y ], Z〉 − 〈[Y, Z], X〉+ 〈[Z,X], Y 〉

=〈[X, Y ], Z〉

∇XY =1

2[X,Y ]

119

Page 127: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

We conveniently write the connection as

ΘL =1

2ad(θ) ∈ Ω1(G; so(g)),

where ad : g → so(g) is the adjoint representation. To calculate the curvature on

left-invariant vector fields X, Y, Z,

RX,YZ =∇X∇YZ −∇Y∇XZ −∇[X,Y ]Z

=1

4[X, [Y, Z]]− 1

4[Y, [X,Z]]− 1

2[[X, Y ], Z]

=1

4[[X, Y ], Z]− 1

2[[X, Y ], Z]

=− 1

4[[X, Y ], Z]

Equivalently, we could use the relation

Ω = dΘ +1

2[Θ ∧Θ].

On vectors X,Y ∈ g

dΘL(X, Y ) = −1

2ad[X, Y ],

and hence dΘL = −14ad[θ ∧ θ]. Therefore,

ΩL = −1

4ad[θ ∧ θ] +

1

8ad[θ ∧ θ] = −1

8ad[θ ∧ θ].

This is equivalent to ΩL(X, Y ) = −14ad[X,Y ]. Therefore,

p∗α(Θ) = 〈ΘL ∧ ΩL〉so(g) −1

6〈ΘL ∧ [ΘL ∧ΘL]〉so(g)

= 〈12ad(θ) ∧ −1

8ad[θ ∧ θ]〉so(g) −

1

6〈12ad(θ) ∧ ad[1

2θ ∧ 1

2θ]〉so(g)

= − 1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g).

Due to the Ad-invariance of the 3-form, this evaluates as

− 1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g)(X, Y, Z) = −1

2〈ad(θ(X)) ∧ ad[θ(Y ), θ(Z)]〉so(g).

120

Page 128: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

By the Ad-invariance of the inner product 〈·, ·〉so(g) and the fact that ad : g→ so(g)

is a representation, the form is bi-invariant and hence harmonic. Therefore, by

Proposition 6.4.5,

HL = − 1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g).

Now, we do the same, but with right-invariant framing. Right-multiplication

gives an isomorphism

TG ∼= G× TeG,

and for right-invariant vector fields X,Y, Z (viewed as elements in TeG under the

above isomorphism) we see that

〈∇XY, Z〉 =1

2〈[X,Y ], Z〉

∇XY =1

2[X, Y ]

RX,YZ = −1

4[[X, Y ], Z].

The bracket is the bracket on vector fields. Let gR be the Lie algebra of right-

invariant vector fields, and let θR : TG → gR be the 1-form that associates to any

vector the associated right-invariant vector field. Under the isomorphism θR : TeG ∼=

gR, we have

ΘR =1

2ad(θR) ∈ Ω1(G; so(gR)),

ΩR = −1

8ad[θR, θR] ∈ Ω2(G; so(gR)),

where ad : gR → so(gR) is the adjoint representation. Therefore,

α(Θ)R = − 1

12〈ad(θR) ∧ [ad(θR) ∧ ad(θR)]〉so(gR) ∈ Ω3(G; so(gR)).

This form is harmonic, and hence equals α(Θ)R. To compare this with α(Θ)L, we

remember the following relationship between left and right-invariant vector fields.

121

Page 129: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

If X ∈ TeG, and XL = θ(X), XR = θR(X) are the associated left and right-invariant

vector fields, then

[XL, YL]|e = −[XR, YR]|e ∈ TeG.

In other words, the two Lie brackets defined on the tangent space TeG differ from

each other by a sign. Consequently, if we consider Xe, Ye, Ze ∈ TeG, as tangent

vectors at a point, XL, YL, ZL ∈ g the induced left-invariant vector fields, and

XR, YR, ZR ∈ gR the induced right-invariant vector fields, we see that

α(Θ)R(Xe, Ye, Ze) = −1

2〈ad(XR), ad[YR, ZR]〉so(gR) = −1

2〈ad(XL),−ad[YL, ZL]〉so(g)

=1

2〈ad(XL), ad[YL, ZL]〉so(g).

Therefore,

α(Θ)R = −α(Θ)L =1

12〈ad(θ) ∧ ad[θ ∧ θ]〉so(g) ∈ Ω3(M).

122

Page 130: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

CHAPTER 7

CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING

STRUCTURES

Suppose we fix a closed compact manifold M along with a choice of spin struc-

ture. Assume that M admits a string structure; i.e. that p1

2(M) = 0 ∈ H4(M ; Z).

Then, given any Riemannian metric g, we can form the Spin(n) frame bundle

Spin(M) = Spin(TM). The Levi-Civita connection induces a canonical connec-

tion on this bundle, which we will denote Θg. By the construction in 6.4.4, we have

a canonical map

Met(M)× string structures on M −→ Ω3(M)

g, S 7−→ Hg,S

The formHg,S is the canonical 3-form given by the equation (based on (M,Spin(M), g,Θg))

limδ→0

[S]gδ= α(Θg)− π∗Hg,S.

Remark 7.0.2. We don’t really have the space Met(M)×string structures on M.

The space of string structures depends on the principal bundle, which varies over the

space of metrics. We really mean points in a fiber bundle over Met(M). However,

since the fibers are discrete and the space of metrics is contractible, choosing a metric

123

Page 131: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

g′ gives a canonical trivialization of the bundle as

Met(M)× string structures on Spin(M, g′)

In general, the space of connections on a principal G-bundle is an affine space

modeled on Ω1(M ; gP ), i.e. any two connections differ by a 1-form with values in

the adjoint bundle. When the bundle P = SO(M) or Spin(M),

Ω1(M ; gP ) = Ω1(M ; so(TM))

where elements of so(TM) are skew-symmetric endomorphisms of the tangent space

at each point. Given a 3-form on H ∈ Ω3(M), the metric can be used to create

an endomorphism-valued 1-form T ∈ Ω1(M ; so(TM)). This follows from the homo-

morphism

Λ3TM∗ −→ TM∗ ⊗ Λ2TM∗ g−→ TM∗ ⊗ so(TM)

induced by the metric g. More explicitly, for any vector field Z ∈ C∞(M,TM),

ιZH ∈ Ω2(M)g∼= so(TM).

Therefore, given Hg,S ∈ Ω3(M), we define T = T g,S ∈ Ω1(M ; so(M)) by

〈T g,S(X, Y ), Z〉 = Hg,S.

Since H is skew-symmetric in all 3-variables, by definition we see that

〈T (X, Y ), Z〉 = 〈T (Y, Z), X〉 = −〈T (Y,X), Z〉.

In other words, the tensor g(T (·, ·), ·) is skew-symmetric in all 3-variables, or totally

skew-symmetric.

124

Page 132: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Given Hg,S ∈ Ω3(M), we can form the connection

Θg,S = Θg +1

2T g,S.

We will be more concerned with the induced vector bundle connection ∇S = ∇g,S

on TM , which we denote

∇S = ∇g +1

2T g,S.

This is the unique metric connection whose torsion T satisfies

〈T (X, Y ), Z〉 = Hg,S(X, Y, Z).

Proposition 7.0.3. A metric connection on TM is completely characterized by its

torsion.

Proof. For X, Y, Z vector fields on M , then

(1) X〈Y, Z〉 = 〈∇XY, Z〉+ 〈Y,∇XZ〉

(2) Y 〈Z,X〉 = 〈∇YZ,X〉+ 〈Z,∇YX〉

(3) Z〈X, Y 〉 = 〈∇ZX, Y 〉+ 〈X,∇ZY 〉

Taking (1) + (2)− (3) and using T (X, Y ) = ∇XY −∇YX − [X, Y ] we get

2〈∇XY, Z〉 = X〈Y, Z〉+ Y 〈Z,X〉 − Z〈X, Y 〉

+ 〈[X, Y ], Z〉 − 〈[Y, Z], X〉+ 〈[Z,X], Y 〉

+ 〈T (X, Y ), Z〉 − 〈T (Y, Z), X〉+ 〈T (Z,X), Y 〉.

Therefore, the connection is completely determined by the metric and torsion tensor.

125

Page 133: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Furthermore, if we assume that T is totally skew-symmetric, then

2〈∇XY, Z〉 = X〈Y, Z〉+ Y 〈X,Z〉 − Z〈X, Y 〉

+ 〈[X, Y ], Z〉 − 〈[Y, Z], X〉+ 〈[Z,X], Y 〉]

+ 〈T (X,Y ), Z〉

〈∇XY, Z〉 = 〈∇gXY +

1

2T (X, Y ), Z〉

In the above construction, the form Hg,S was constructed as a harmonic repre-

sentative in a scaling limit, and hence the form is independent of the global scaling

on M . However, we use the metric to transform Hg,S into an endomorphism-valued

1-form. Therefore, the torsion tensor we create depends on a global scaling. In fact,

if we let

g′ = εg, ε ≥ 0,

then the torsion tensor T g,S transforms as

ε〈T εg,S(X, Y ), Z〉g = 〈T εg,S(X, Y ), Z〉εg = Hg,S(X, Y, Z) = 〈T g,S(X, Y ), Z〉g.

Therefore,

T εg,S =1

εT g,S.

Consequently, as ε → ∞, or the manifold becomes scaled very large, the torsion

tensor goes to 0. In the opposite scaling, as M becomes very small, the torsion

tensor (unless it is zero), becomes very large and does not even converge to an

element of Ω1(M ; so(TM)).

Combining the above, for any Riemannian manifold (M, g) with string structure

S ∈ H3(Spin(M); Z), we have a canonical 1-parameter family of metric connections

126

Page 134: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

given by

Θεg,S = Θεg +1

2T εg,S,

〈∇εg,SX Y, Z〉εg = 〈∇εg

XY, Z〉εg +1

2Hg,S(X, Y, Z).

Since the connection∇g is invariant under rescaling, it is more convenient to consider

this as a 1-paramter family of connections on the fixed metric g. In other words

〈∇εg,SX Y, Z〉g = 〈∇g

XY, Z〉g +1

2εHg,S(X, Y, Z).

At this point, we will deal with a fixed metric and string structure. The correspond-

ing notation will be dropped when it is unambiguous. In particular, T = T g,S will be

the torsion tensor of the modified connection, and, to avoid excessive parenthesis,

we denote

TXY = T (X, Y ).

Proposition 7.0.4. For the connection ∇S = ∇g + 12T , where T = T g,S,

RicS(X, Y ) = Ricg(X, Y ) +1

2

∑i

〈(∇geiT )(X, Y ), ei〉 −

1

4

∑i

〈TeiX,Tei

Y 〉.

For ∇εg,S = ∇g + 12εT ,

Ricεg,S(X, Y ) = Ricg(X, Y ) +1

∑i

〈(∇geiT )(X, Y ), ei〉 −

1

4ε2

∑i

〈TeiX,Tei

Y 〉.

127

Page 135: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

Proof. Let R denote the curvature tensor. Then,

RSX,YZ =∇S

X∇SYZ −∇S

Y∇SXZ −∇S

[X,Y ]Z

=(∇gX +

1

2TX)(∇g

Y +1

2TY )Z − (∇g

Y +1

2TY )(∇g

X +1

2TX)Z − (∇g

[X,Y ] +1

2T[X,Y ])Z

=(∇g

X∇gYZ −∇

gY∇

gXZ −∇

g[X,Y ]Z

)+

1

2(∇g

XTYZ − TY∇gXZ)

− 1

2(∇g

Y TXZ − TX∇gYZ) +

1

4(TXTYZ − TY TXZ)− 1

2T[X,Y ]Z

=RgX,YZ +

1

2

((∇g

XT )(Y, Z) + T∇gXYZ

)− 1

2

((∇g

Y T )(X,Z) + T∇gY XZ

)+

1

4(TXTYZ − TY TXZ)− 1

2T[X,Y ]Z

Let ei be an orthonormal frame. The Ricci tensor, denoted Ric, is defined by

Ric(Y, Z) =n∑

i=1

〈Rei,YZ, ei〉.

Using the fact that 〈TXY, Z〉 is skew-symmetric in all 3-variables, we see

〈RSei,Y

Z, ei〉 =〈Rgei,Y

Z, ei〉+1

2〈(∇g

eiT )(Y, Z), ei〉+

1

2〈T∇g

eiYZ, ei〉 −

1

2〈(∇g

Y T )(ei, Z), ei〉

− 1

2〈T∇g

Y eiZ, ei〉+

1

4〈Tei

TYZ, ei〉 −1

4〈TY Tei

Z, ei〉 −1

2〈T[ei,Y ]Z, ei〉

=〈Rgei,Y

Z, ei〉+1

2〈(∇g

eiT )(Y, Z), ei〉+

1

2〈T∇g

eiYZ − T∇g

Y eiZ − T[ei,Y ]Z, ei〉

− 1

2〈(∇g

Y T )(ei, Z), ei〉 −1

4〈TY Tei

Z, ei〉

=〈Rgei,Y

Z, ei〉+1

2〈(∇g

eiT )(Y, Z), ei〉 −

1

4〈Tei

Y, TeiZ〉

The term 14〈Tei

TYZ, ei〉 drops out in the second equality due to the skew-symmetry

of 〈T (·, ·), ·〉. In the last equality, we have that

1

2〈T∇g

eiYZ − T∇g

Y eiZ − T[ei,Y ]Z, ei〉 =

1

2〈TT g

eiYZ, ei〉 = 0

128

Page 136: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

due to the torsion-freeness of the Levi-Civita connection. Also,

〈(∇geiT )(Y, Z), ei〉 =〈∇g

Y TeiZ, ei〉 − 〈T∇g

Y eiZ, ei〉 − 〈Tei

∇gYZ, ei〉

=〈∇gY Tei

Z, ei〉+ 〈TeiZ,∇g

Y ei〉

=Y 〈TeiZ, ei〉 = 0

Therefore, we have the relationship

RicS(X, Y ) = Ricg(X, Y ) +1

2

∑i

〈(∇geiT )(X, Y ), ei〉 −

1

4

∑i

〈TeiX,Tei

Y 〉.

If one keeps track of the ε in the above calculation for ∇εg,S = ∇g + 12εT , then

Ricεg,S(X, Y ) = Ricg(X, Y ) +1

∑i

〈(∇geiT )(X, Y ), ei〉 −

1

4ε2

∑i

〈TeiX,Tei

Y 〉.

The modified Ricci tensor contains the original Ricci tensor, plus two other

terms. The first is skew-symmetric, and the second is symmetric. In particular, if

we look at Ric(X) = Ric(X,X), then

Ricεg,S(X) = Ricg(X)− 1

4ε2

∑i

‖TeiX‖2. (7.0.1)

Consequently, if HS 6= 0, then for some X,

Ricεg,S(X)ε→0−→ −∞.

Conversely, if the 1-parameter family of modified connections have positive Ricci

curvature for ε close to 0 (i.e. if Ricεg,S > 0 for ε ∼ 0), then Ricε,S > 0 for all ε and

HS = 0. In this situation, the 1-parameter family of metrics is really the Levi-Civita

connection for a metric of positive Ricci curvature.

Remark 7.0.5. It was originally hoped that given (M, g, S), the canonical metric

connection ∇g,S would have its modified Pontryagin form p1

2(∇g,S) = 0. The above

129

Page 137: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

discussion shows this can not be true, since the Pontryagin form from the Levi-Civita

connection is invariant under a global rescaling, and the connection ∇εg,S converges

to the Levi-Civita connection as ε → ∞. Therefore, if p1

2(M, g) 6= 0 ∈ Ω4(M), the

modified Pontryagin form can not always be zero. One might hope that in the other

scaling limit, as ε → 0, the modified Pontryagin form is zero. However, one can

show this is not the case either. This can be seen by a brute force construction on

S1 × SU(2), where the metric is defined as a 1-parameter family of left-invariant

metrics on SU(2).

130

Page 138: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

BIBLIOGRAPHY

[AHS] M. Ando, M. J. Hopkins, and N. P. Strickland. Elliptic spectra, the Wittengenus and the theorem of the cube. Invent. Math., 146(3):595–687, 2001.

[AS1] Orlando Alvarez and I. M. Singer. Beyond the elliptic genus. Nuclear Phys.B, 633(3):309–344, 2002.

[AS2] M. F. Atiyah and L. Smith. Compact Lie groups and the stable homotopyof spheres. Topology, 13:135–142, 1974.

[BF] Jean-Michel Bismut and Daniel S. Freed. The analysis of elliptic families. II.Dirac operators, eta invariants, and the holonomy theorem. Comm. Math.Phys., 107(1):103–163, 1986.

[CE] Claude Chevalley and Samuel Eilenberg. Cohomology theory of Lie groupsand Lie algebras. Trans. Amer. Math. Soc., 63:85–124, 1948.

[CP] R. Coquereaux and K. Pilch. String structures on loop bundles. Comm.Math. Phys., 120(3):353–378, 1989.

[CS] Shiing Shen Chern and James Simons. Characteristic forms and geometricinvariants. Ann. of Math. (2), 99:48–69, 1974.

[Dai] Xianzhe Dai. Adiabatic limits, nonmultiplicativity of signature, and Lerayspectral sequence. J. Amer. Math. Soc., 4(2):265–321, 1991.

[For] Robin Forman. Spectral sequences and adiabatic limits. Comm. Math. Phys.,168(1):57–116, 1995.

[Fre1] Daniel S. Freed. Classical Chern-Simons theory. I. Adv. Math., 113(2):237–303, 1995.

[Fre2] Daniel S. Freed. Classical Chern-Simons theory. II. Houston J. Math.,28(2):293–310, 2002.

[Hit] Nigel Hitchin. Harmonic spinors. Advances in Math., 14:1–55, 1974.

[Hop] M. J. Hopkins. Algebraic topology and modular forms. In Proceedings ofthe International Congress of Mathematicians, Vol. I (Beijing, 2002), pages291–317, Beijing, 2002. Higher Ed. Press.

[KN] Shoshichi Kobayashi and Katsumi Nomizu. Foundations of differential ge-ometry. Vol I. Interscience Publishers, a division of John Wiley & Sons,New York-Lond on, 1963.

131

Page 139: CANONICAL METRIC CONNECTIONS ASSOCIATED TO STRING STRUCTURES A

[Lic] Andre Lichnerowicz. Laplacien sur une variete riemannienne et spineurs.Atti Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Nat. (8), 33:187–191, 1962.

[LM] H. Blaine Lawson, Jr. and Marie-Louise Michelsohn. Spin geometry, vol-ume 38 of Princeton Mathematical Series. Princeton University Press,Princeton, NJ, 1989.

[MM] Rafe R. Mazzeo and Richard B. Melrose. The adiabatic limit, Hodge coho-mology and Leray’s spectral sequence for a fibration. J. Differential Geom.,31(1):185–213, 1990.

[Seg1] Graeme Segal. Elliptic cohomology (after Landweber-Stong, Ochanine, Wit-ten, and others). Asterisque, (161-162):Exp. No. 695, 4, 187–201 (1989),1988.

[Seg2] Graeme Segal. The definition of conformal field theory. In Topology, geometryand quantum field theory, volume 308 of London Math. Soc. Lecture NoteSer., pages 421–577. Cambridge Univ. Press, Cambridge, 2004.

[ST] Stephan Stolz and Peter Teichner. What is an elliptic object? In Topology,geometry and quantum field theory, volume 308 of London Math. Soc. LectureNote Ser., pages 247–343. Cambridge Univ. Press, Cambridge, 2004.

[Sto] Stephan Stolz. A conjecture concerning positive Ricci curvature and theWitten genus. Math. Ann., 304(4):785–800, 1996.

[Wit1] Edward Witten. Global gravitational anomalies. Comm. Math. Phys.,100(2):197–229, 1985.

[Wit2] Edward Witten. The index of the Dirac operator in loop space. In Ellip-tic curves and modular forms in algebraic topology (Princeton, NJ, 1986),volume 1326 of Lecture Notes in Math., pages 161–181. Springer, Berlin,1988.

[Wit3] Edward Witten. Index of Dirac operators. In Quantum fields and strings:a course for mathematicians, Vol. 1, 2 (Princeton, NJ, 1996/1997), pages475–511. Amer. Math. Soc., Providence, RI, 1999.

132