Math & Physics: Bridging Concepts
Math & Physics: Bridging Concepts
Volume 141
Editors
S.S Antman
Department of Mathematics
and
Institute for Physical Science and Technology
University of Maryland
College Park, MD 20742-4015
USA
[email protected]
P. Holmes
Department of Mechanical and Aerospace Engineering
Princeton University
215 Fine Hall
Princeton, NJ 08544
[email protected]
L. Sirovich
Laboratory of Applied Mathematics
Department of Biomathematical Sciences
Mount Sinai School of Medicine
New York, NY 10029-6574
[email protected]
K. Sreenivasan
Department of Physics
New York University
70 Washington Square South
New York City, NY 10012
[email protected]
Advisors
L. Greengard J. Keener
J. Keller R. Laubenbacher B.J. Matkowsky
A. Mielke C.S. Peskin A. Stevens A. Stuart
Topology, Geometry
and Gauge fields
Interactions
Second Edition
Gregory L. Naber
Drexel University
Department of Mathematics
Korman Center
3141 Chestnut Street
Philadelphia, Pennsylvania 19104-2875
USA
[email protected]
ISSN 0066-5452
ISBN 978-1-4419-7894-3 e-ISBN 978-1-4419-7895-0
DOI 10.1007/978-1-4419-7895-0
Springer New York Dordrecht Heidelberg London
Library of Congress Control Number: 2011923654
vii
viii Preface
ix
Contents
Chapter 1
Geometrical Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Smooth Manifolds and Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Matrix Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Principal Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.4 Connections and Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.5 Associated Bundles and Matter Fields . . . . . . . . . . . . . . . . . . . . . . 38
Chapter 2
Physical Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.1 General Framework for Classical Gauge Theories . . . . . . . . . . . . 45
2.2 Electromagnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.3 Spin Zero Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.4 Spin One-Half Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2.5 SU (2)-Yang-Mills-Higgs Theory on Rn . . . . . . . . . . . . . . . . . . . . . 104
2.6 Epilogue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Chapter 3
Frame Bundles and Spacetimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.1 Partitions of Unity, Riemannian Metrics and Connections . . . . . 139
3.2 Continuous Versus Smooth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
3.3 Frame Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
3.4 Minkowski Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
3.5 Spacetime Manifolds and Spinor Structures . . . . . . . . . . . . . . . . . 168
Chapter 4
Differential Forms and Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
4.1 Multilinear Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
4.2 Vector-Valued Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
4.3 Differential Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
4.4 The de Rham Complex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
4.5 Tensorial Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.6 Integration on Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.7 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Chapter 5
de Rham Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
5.1 The de Rham Cohomology Groups . . . . . . . . . . . . . . . . . . . . . . . . 258
5.2 Induced Homomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
5.3 Cochain Complexes and Their Cohomology . . . . . . . . . . . . . . . . . 275
5.4 The Mayer-Vietoris Sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
xi
xii Contents
Chapter 6
Characteristic Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
6.2 Algebraic Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
6.3 The Chern-Weil Homomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . 317
6.4 Chern Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
6.5 Z2 -Čech Cohomology for Smooth Manifolds . . . . . . . . . . . . . . . . . 335
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Seiberg-Witten Gauge Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
A.1 Donaldson Invariants and TQFT . . . . . . . . . . . . . . . . . . . . . . . . . . 351
A.2 Clifford Algebra and Spinc -Structures . . . . . . . . . . . . . . . . . . . . . . 363
A.3 Seiberg-Witten Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
A.4 The Moduli Space and Invariant . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
A.5 The Witten Conjecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413
1
Geometrical Background
real line R with its standard differentiable structure (determined by the at-
las consisting of the single chart (R, id)) we denote by C ∞ (X) the set of all
smooth, real-valued functions on X and provide it with the obvious (point-
wise) structure of a commutative algebra with identity.
A tangent vector at p ∈ X is a real-valued function v : C ∞ (X) −→ R
that is linear and satisfies the Leibnitz Product Rule v (f g) = f (p)v (g) +
v (f )g(p) for all f, g ∈ C ∞ (X). The collection of all such is denoted Tp (X),
called the tangent space to X at p and provided with the natural point-
wise structure of a real vector space. The dimension of Tp (X) as a vector
space over R is the same as the dimension of X as a manifold. Indeed, if
(U, φ) is a chart at p in X with coordinate functions xi , i = 1, . . . , n (φ(p) =
(x1 (p), . . . , xn (p))), then the linear maps ∂i |p = ∂x ∂
i |p : C
∞
(X) −→ R defined
−1
by ∂xi |p (f ) = Di (f ◦ φ )(φ(p)) are in Tp (X) and { ∂x1 |p , . . . , ∂x∂n |p } is a ba-
∂ ∂
sis for Tp (X) (Theorem 5.5.3, [N4]). Any v ∈ Tp (X) can be uniquely written
i |p (summation convention). If (a, b) is an open interval in R
∂
as v = v (xi ) ∂x
provided with its standard differentiable structure (determined by the atlas
consisting of the single chart ((a, b), ι), where ι : (a, b) ,→ R is the inclusion
map), then a smooth map α : (a, b) −→ X is a smooth curve in X. Fix
t0 ∈ (a, b) and let p = α(t0 ). The velocity vector of α at t0 is the map
α′ (t0 ) : C ∞ (X) −→ R defined by (α′ (t0 ))(f ) = D1 (f ◦ α)(t0 ) for each f in
C ∞ (X). Then α′ (t0 ) is in Tp (X) and, indeed, every element of Tp (X) is the
velocity vector of some smooth curve in X through p (Corollary 5.5.6, [N4]).
If f : X −→ Y is a smooth map and p ∈ X, then the derivative of
f at p is the linear map f∗p : Tp (X) −→ Tf (p) (Y ) defined as follows: For
each v ∈ Tp (X), f∗p (v ) : C ∞ (Y ) −→ R is given by (f∗p (v ))(g) = v (g ◦ f )
for all g ∈ C ∞ (Y ). If v = α′ (t0 ) for some smooth curve α, then f∗p (v ) =
f∗p (α′ (t0 )) = (f ◦α)′ (t0 ). f is said to be an immersion at p if f∗p is one-to-one
and an immersion if this is true at each p ∈ X. f is a submersion at p if
f∗p is onto and a submersion if this is true at each p ∈ X. An immersion
that is also a homeomorphism onto its image is an imbedding. A point q ∈ Y
is a regular value of f if, for every p ∈ f −1 (q), f is a submersion at p (this
is the case, in particular, if f −1 (q) = ∅); otherwise, q is a critical value of f .
If X ′ is an open subspace of X and {(Uα , φα )}α∈A is an atlas for X, then
{(Uα ∩ X ′ , φα |Uα ∩ X ′ ) : α ∈ A, Uα ∩ X ′ ̸= ∅} is an atlas for X ′ and, with
the differentiable structure determined by this atlas, X ′ is called an open
submanifold of X. Note that dim X ′ = dim X. More generally, if dim X =
n and 1 ≤ k ≤ n is an integer, then a topological subspace X ′ of X is
called a k -dimensional submanifold of X if, for each p ∈ X ′ , there exists
a chart (U, φ) in the differentiable structure for X such that φ(U ∩ X ′ ) =
{(x1 , . . . , xk , xk+1 , . . . , xn ) ∈ φ(U ) : xk+1 = · · · = xn = 0}. For each such
(U, φ) one obtains a chart (U ∩ X ′ , φ′ ) for X ′ , where φ′ is φ|U ∩ X ′ followed
by the projection of Rn = Rk × Rn−k onto Rk . The collection of all such (U ∩
X ′ , φ′ ) is an atlas for X ′ and so determines a differentiable structure for X ′ . A
0-dimensional submanifold of X is a discrete subspace of X. Restrictions
of C ∞ maps on X to submanifolds are C ∞ with respect to this submanifold
1.1. Smooth Manifolds and Maps 3
φ−1 −1 1
N (y) = φN (y , . . . , y ) = (1 + ∥y∥ )
n 2 −1
(2y 1 , . . . , 2y n , −∥y∥2 + 1). Thus, on
φN (UN ∩US ) = φS (UN ∩US ) = R −{0}, φS ◦φ−1
n −1
N (y) = φN ◦φS (y) = ∥y∥
−2
y
∞
and these are C . Thus, {(US , φS ), (UN , φN )} is an atlas for S and so de- n
φ2 ◦ φ−1 2 3 n 2 3 n 2 −1 3 2 −1
1 (y , y , . . . , y ) = φ2 ([1, y , y , . . . , y ]) = ((y ) , y (y ) , . . . ,
y (y ) ). Thus {(Uk , φk )}k=1,...,n is an atlas for FP
n 2 −1 n−1
and so determines a
differentiable structure.
Remark: When n = 2 we have RP1 ∼
= S 1 , CP1 ∼
= S 2 and HP1 ∼
= S4
(pages 53–54, [N4]).
6. (Classical Groups) Let F = R, C, or H and let GL(n, F) be the set of
all n × n matrices with entries in F (the reason for the peculiar notation will
emerge in Section 1.2). Topologically we identify GL(n, R) = Rn , GL(n, C) =
2
R2n and GL(n, H) = R4n . Let GL(n, F) denote the set of all invertible ele-
2 2
[W , V ] = −[V , W ]
[a1 V 1 + a2 V 2 , W ] = a1 [V 1 , W ] + a2 [V 2 , W ]
[f V , gW ] = f g[V , W ] + f (V g)W − g(W f )V
[V 1 , [V 2 , V 3 ]] + [V 3 , [V 1 , V 2 ]] + [V 2 , [V 3 , V 1 ]] = 0
( i i
)
j ∂W j ∂V ∂
[V , W ] = V j
−W j
∂x ∂x ∂xi
∂F i
F ∗Θ = (Θi ◦ F )dxj
∂xj
= Θi (F 1 (x1 , . . . , xn ), . . . , F m (x1 , . . . , xn ))d(F i (x1 , . . . , xn )),
∂f i
df = dx (f ∈ C ∞ (X)).
∂xi
If X ′ is a submanifold of X and ι : X ′ ,→ X is the inclusion map, then the
restriction of Θ ∈ X ∗ (X) to X ′ is defined to be ι∗ Θ. Finally, if F : X −→ Y
and G : Y −→ Z are smooth, then
(G ◦ F )∗ = F ∗ ◦ G∗
A = Aij ei ⊗ ej = A (ei , ej ) ei ⊗ ej
1.1. Smooth Manifolds and Maps 9
ω ∧ρ η = (ω 1 + ω 2 i ) ∧ρ (η 1 + η 2 i )
= (ω 1 ∧ η 1 − ω 2 ∧ η 2 ) + (ω 1 ∧ η 2 + ω 2 ∧ η 1 )i .
6. Any (finite) product of Lie groups (with the product manifold structure
and the direct product group structure) is a Lie group, e.g., SU (2) × U (1),
SU (3) ×SU (2) × U (1), or any torus S 1 × · · · × S 1 .
Two Lie groups G1 and G2 are isomorphic if there is a diffeomorphism of
G1 onto G2 that is also a group isomorphism, e.g., S 1 is isomorphic to both
U (1) and SO(2), while S 3 is isomorphic to SU (2) and Sp(1). Any subgroup of
some GL(n, C) that is also a submanifold is called a matrix Lie group and
we shall henceforth restrict our attention to these. GL(n, R) can be identified
with a subgroup of GL(n, C) that is also a submanifold. It is less obvious, but
also true, that GL(n, H) can be identified with a subgroup of GL(2n, C) that
is also a submanifold. To see this observe first that any quaternion x0 + x1 i +
x2 j + x3 k can be written as z 1 + z 2 j , where z 1 = x0 + x1 i and z 2 = x2 + x3 i ,
and so identified with a pair of complex numbers. Thus, any n×n quaternionic
matrix P can be written as P = A + Bj , where A and B are n × n complex
matrices. Define a map ϕ : GL (n, H) −→ GL(2n, C) by
( )
A B
ϕ(P ) = .
−B̄ Ā
Then ϕ is an algebra isomorphism that also preserves the conjugate transpose
(Exercise 1.1.28, [N4]). In particular, P ∈ Sp(n) if and only
( if ϕ(P) ) ∈ U (2n).
Moreover, a 2n × 2n complex matrix M has the form −B̄ A B
Ā
if and only
−1
( 0 id )
if it satisfies JM J = M̄ , where J = −id 0 and, if M is unitary, this is
T
equivalent to M JM = J. Thus, we may identify
The real and complex special linear groups SL(n, R) and SL(n, C) are
the subgroups of GL(n, R) and GL(n, C), respectively, consisting of those el-
ements with determinant 1. Although the noncommutativity of H blocks any
obvious notion of a determinant for quaternionic matrices, one can define the
quaternionic special linear group SL(n, H ) to be the subset of GL(n, H)
consisting of all those elements P such that det ϕ(P ) = 1. This is, indeed,
a subgroup of GL(n, H) (Exercise 1.1.30, [N4]). One can show directly that
GL(n, H), Sp(n) and SL(n, H) are all submanifolds of GL(2n, C), but there
is also a general result to the effect that any closed subgroup of a complex
general linear group is necessarily a submanifold (see [Howe] or, for a still
more general result, [Warn]). We conclude then that all of the classical groups
are matrix Lie groups.
A Lie algebra is a real vector space L on which is defined a bilinear opera-
tion [ , ] : L×L −→ L, called bracket, that is skew-symmetric ([y, x] = −[x, y]
for all x, y ∈ L) and satisfies the Jacobi identity ([x, [y, z]] + [z, [x, y]] +
[y, [z, x]] = 0 for all x, y, z ∈ L). Two Lie algebras L1 and L2 with brackets
1.2. Matrix Lie Groups 13
For matrix Lie groups G, Tid (G) can be identified with a set of matrices
(velocity vectors at t = 0 to smooth curves t −→ (aij (t)) with (aij (0)) = id,
computed entrywise). One can show (Section 5.8, [N4]) that this set of
matrices is always closed under the formation of commutators and that the
14 1. Geometrical Background
Isomorphic Lie groups have isomorphic Lie algebras. For example, since
U (1) and SO(2) are isomorphic, so are u(1) and so(2) and, since u(1) is the
algebra of 1 × 1 skew-Hermitian matrices, both can be naturally identified
with the Lie algebra Im C. Similarly, SU (2) and Sp(1) are isomorphic and
therefore so are their Lie algebras su(2) and sp(1). But sp(1) is naturally
identfied with the Lie algebra Im H of pure imaginary quaternions which,
in turn, is isomorphic to R3 with the cross product as bracket. However,
non-isomorphic Lie groups can have isomorphic Lie algebras, e.g., O(n) and
SO(n) in #3 above. A less trivial example consists of SU (2) and SO(3).
These are certainly not isomorphic as Lie groups. Indeed, they are not even
homeomorphic since SU (2) ∼ = S 3 (Theorem 1.1.4, [N4]) and S 3 is simply
connected (page 119, [N4]), whereas SO(3) ∼ = RP3 (page 399, [N4]) and
π1 (R P3 ) ∼ = 2 Z (Theorem 2.4.5, [N4]). To see that the Lie algebras su(2)
and so(3) are isomorphic we introduce the notion of a Lie group’s “structure
constants.” Let G be an arbitrary Lie algebra. Select a basis {e1 , . . . , en } for
G. For all i, j = 1, . . . , n, [ei , ej ] is in G so there exist unique constants Cij k
,
k k
k = 1, . . . , n, such that [ei , ej ] = Cij ek . The constants Cij , i, j, k = 1, . . . , n are
called the structure constants of G relative to the basis {e1 , . . . , en }. Two
1.2. Matrix Lie Groups 15
Lie algebras are clearly isomorphic if and only if there exist bases relative
to which the structure constants are the same. Now, an obvious basis for
so(3) (3 × 3 real, skew-symmetric matrices) consists of
0 0 0 0 0 1 0 −1 0
τ1 =
0 0 −1 , τ2 = 0 0 0 , τ3 = 1
0 0
0 1 0 −1 0 0 0 0 0
∑3
and a simple calculation shows that [τi , τj ] = k=1 ϵijk τk , where ϵijk is the
Levi-Civita symbol (1 if ijk is an even permutation of 123, −1 if ijk is an
odd permutation of 123, and 0 otherwise). The usual basis for su(2) consists
of Tk = − 12 i σk , where
( ) ( ) ( )
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 =
1 0 i 0 0 −1
are the Pauli spin matrices (page 396, [N4]) and here again one quickly
∑3
verifies that [Ti , Tj ] = k=1 ϵijk Tk . Thus, so(3) and su(2)
∑∞ are isomorphic.
For any A ∈ GL(n, C) we define exp(A) = eA = k=0 k! 1 k
A . The series
converges absolutely and uniformly on every bounded region in GL(n, C) =
Cn = R2n . Every exp(A) is invertible (because det(eA ) = etrace(A) ) so exp
2 2
is a C ∞ map of GL(n, C) to GL(n, C). If G is any matrix group, then its Lie
algebra G can be identified with a subalgebra of GL(n, C) for some n and the
restriction of exp to G maps into G (see Theorem 5.8.6, [N4], for the cases
of interest here). The map exp : G −→ G has a derivative at 0 ∈ G and, if
T0 (G) is identified with G by the canonical isomorphism, exp∗o : G −→ G is the
identity map (Lemma 5.8.5 (5), [N4]). It follows from the Inverse Function
Theorem that exp is a diffeomorphism of some neighborhood of 0 ∈ G onto a
neighborhood of exp(0) = id in G, i.e., on some neighborhood of id in G, exp−1
is a chart for G.
A 1-form Θ on a Lie group G (real, or vector-valued) is said to be
left invariant if (Lg )∗ Θ = Θ for all g ∈ G, i.e., if, for all g, h ∈ G,
Θ(h) = (Lg )∗ (Θ(gh)), or, equivalently, Θ(gh) = (Lg−1 )∗ (Θ(h)). This is the
case if and only if Θ(g) = (Lg−1 )∗ (Θ(id)) for every g ∈ G. Any such 1-form is
necessarily smooth (page 285, [N4]) and is uniquely determined by its value
at the identity. In particular, any covector at id in G uniquely determines
a left invariant 1-form taking that value at id. A particularly important ex-
ample of a left invariant 1-form on G is the Cartan (canonical) 1-form
which takes values in the vector space G and is defined as follows: For each
g ∈ G, Θ(g) = Θ g : Tg (G) −→ G = Tid (G) is given by
Θ(g)(v ) = Θ g (v ) = (Lg−1 )∗g (v ).
16 1. Geometrical Background
Θ H = ι∗ Θ G
For calculations
( ∑nit is ik
most convenient
) to identify Θ(g) with the matrix of
−1
components k=1 x (g )dxkj (g) i,j=1,...,n . Note that this is the formal
matrix product of g −1 and
dx11 (g) ... dx1n (g)
.. ..
dx(g) = . .
dxn1 (g) ... dxnn (g)
Since α = x11 (g), β = x12 (g), γ = x21 (g) and δ = x22 (g) we may regard this
as the value at g of the matrix of ordinary 1-forms given by
Θ GL(2,R) =
( )
12 21 −1 x22 dx11 − x12 dx21 x22 dx12 − x12 dx22
11 22
(x x −x x )
−x21 dx11 + x11 dx21 −x21 dx12 + x11 dx22
one finds that, for any g ∈ G, Θ(g) can be identified with g −1 dz(g), where dz
is the matrix of complex-valued 1-forms
dz 11 ... dz 1n dx11 + i dy 11 ... dx1n + i dy 1n
. .. .. ..
dz =
.
.
. = . .
n1 nn n1 n1 nn nn
dz ... dz dx + i dy ... dx + i dy
(complex-valued 1-forms are just vector-valued 1-forms with values in C = R2
and we have used the basis {1, i } for C over R). Thus, if v ∈ Tg (G) (thought
of as an n×n matrix of complex numbers obtained by differentiating entrywise
a smooth curve in GL(n, C) through g), then dz ij picks out its ij-entry v ij
and Θ(g)(v) = g −1 (v ij )i,j=1,...,n . For n = 2 one can write Θ as a matrix of
complex-valued 1-forms just as we did for GL(2, R) in #1:
Θ GL(2,C) =
( )
z 22 dz 11 − z 12 dz 21 z 22 dz 12 − z 12 dz 22
(z 11 z 22 − z 12 z 21 )−1
−z 21 dz 11 + z 11 dz 21 −z 21 dz 12 + z 11 dz 22
18 1. Geometrical Background
3. G = GL(n, H), G = GL(n, H): The result is the same as for GL(n, R) and
GL(n, C) in #1 and #2: For every g ∈ GL(n, H),
Θ(g) = g −1 dq(g),
where
dq 11 ... dq 1n
. ..
dq =
.
. .
dq n1 ... dq nn
and dq ij = dxij + i dy ij + j duij + k dv ij , i, j = 1, . . . , n. This time, however,
there is no simple formula for g −1 when n = 2 since we lack a determinant
function for quaternionic matrices. However, when n = 1 and we identify
a 1 × 1 quaternionic matrix g = (q) with its sole entry q, GL(1, H) is just
the group H − {0} of nonzero quaternions and g −1 dq is just the quaternion
product
1
g −1 dq = q −1 dq = 2 q̄dq.
|q|
Identifying tangent vectors v ∈ Tq (H − {0}) with quaternions v ∈ H via the
canonical isomorphism, q −1 dq(v ) = |q|1 2 q̄v.
4. G = SO(2) ∼ = U (1), G = so(2) ∼
= u(1) ∼= Im C: First regard G as SO(2) ⊆
GL(2, R). Then Θ SO(2) is the restriction of Θ GL(2,R) from #1 to SO(2). One
finds (page 294, [N4]) that
( )
0 x22 dx12 − x12 dx22
Θ SO(2) =
−x22 dx12 + x12 dx22 0
( )
0 −1
= (−x22 dx12 + x12 dx22 ) .
1 0
Now,
( cos θ the)standard identification of SO(2) and U (1) is
−sin θ
−→ (ei θ ) and this induces (by differentiation at id) the
sin θ cos θ ( )
identification θ0 01 −10 −→ θ0 i of so(2) and u(1) ∼
= Im C. Thus,
adg : G −→ G.
Thus, adg = (Adg )∗id = (Lg )∗g−1 ◦ (Rg−1 )∗id = (Rg−1 )∗g ◦ (Lg )∗id and one can
show (Lemma 5.8.7, [N4]) that
adg (A) = g Ag −1
for any A ∈ G. The assignment g −→ adg is a homomorphism of G into GL(G)
and is called the adjoint representation of G on G. The significance of the
adjoint representation may not be immediately apparent, but will become
clear when we discuss connections on principal bundles. On an even more
concrete level, it is shown in Appendix A to [N4] that SO (3) can be identified
20 1. Geometrical Background
with SU (2)/ ± 1 and that, when su(2) is identified with R3 , the adjoint
representation of SU (2) on su (2) is just the natural representation of SO(3)
on R3 by rotation.
We outline a few general facts about Killing forms, although we will require
only special cases (all that we say is proved in [Helg]). If G is a matrix
Lie group with Lie algebra G and if A and B are two fixed elements of G,
then we can define a linear transformation KAB : G −→ G by KAB (X) =
[A, [B, X]] for all X ∈ G. Then the trace of this linear transformation is a
real number (because G is a real vector space) and the map K : G −→ R
defined by K(A, B) = trace (KAB ) is a symmetric, bilinear form on G called
the Killing form of G. K is ad (G)-invariant in the sense that, for any
g ∈ G, K(adg (A), adg (B)) = K(A, B) for all A, B ∈ G. The Lie group G
(or its Lie algebra G) is said to be semisimple if the Killing form K is
nondegenerate.
Remark: There is an algebraic characterization of semisimplicity due to
Cartan. An ideal in a Lie algebra G is a linear subspace H of G with the
property that [A, B] ∈ H whenever A ∈ H and B ∈ G. In particular, an ideal
is itself a Lie algebra under the same bracket operation. A proper ideal is
one that is neither the zero subspace nor the entire Lie algebra. Then G is
semisimple if and only if it can be written as a direct sum of ideals, each of
which (as a Lie algebra) has no proper ideals.
If G is connected and semisimple, then, by a theorem of Weyl, the Killing
form K is negative definite if and only if G is compact. In this case,
one obtains a positive definite, ad(G)-invariant inner product ⟨ , ⟩ on G by
setting ⟨A, B⟩ = −K(A, B). A Riemannian metric g on G is then ob-
tained by left translation, i.e., by defining, for each a ∈ G and all v , w ∈
Ta (G), g a (v , w ) = ⟨(La−1 )∗a (v ), (La−1 )∗a (w )⟩. This metric is left invari-
ant, i.e., L∗a g = g for all a ∈ G, because (L∗a g )b (v , w ) = g ab ((La )∗b (v ),
(La )∗b (w )) = ⟨(L(ab)−1 )∗ab ((La )∗b (v )), (L(ab)−1 )∗ab ((La )∗b (w ))⟩ = ⟨(Lb−1 )∗b
(v ), (Lb−1 )∗b (w )⟩ = g b (v , w ). It is also right invariant, i.e., Ra∗ g = g for
all a ∈ G, because
A metric on a Lie group that is both left invariant and right invariant is
said to be bi-invariant and we have just shown that the Killing form on
1.2. Matrix Lie Groups 21
of su(2) with Im H ∼
= R3 , ⟨ , ⟩ is just twice the usual inner product on R3 .
Next suppose that G is a Lie group and P is a differentiable manifold. A
smooth right action of G on P is a C ∞ map σ : P ×G −→ P which satisfies
1. σ(p, e) = p for all p ∈ P (e is the identity element in G), and
2. σ(p, g1 g2 ) = σ(σ(p, g1 ), g2 ) for all g1 , g2 ∈ G and all p ∈ P .
One generally writes σ(p, g) = p · g and thinks of g as “acting on” p to produce
p · g ∈ P . Then the defining properties assume the form
1. p · e = p for all p ∈ P , and
2. p · (g1 g2 ) = (p · g1 ) · g2 for all g1 , g2 ∈ G and all p ∈ P .
For each fixed g ∈ G we define σg : P −→ P by σg (p) = p · g. Then σg is
a diffeomorphism of P onto P with inverse σg−1 . Similarly, a smooth left
action of G on P is a smooth map ρ : G × P −→ P, ρ(g, p) = g · p, that
satisfies
1. e · p = p for all p ∈ P , and
2. (g1 g2 ) · p = g1 · (g2 · p) for all g1 , g2 ∈ G and all p ∈ P .
The maps ρg : P −→ P defined by ρg (p) = g · p are all diffeomorphisms.
Remark: If (g, p) −→ g · p is a left action, then (p, g) −→ p ⊙ g = g −1 · p
is a right action and, if (p, g) −→ p · g is a right action, then (g, p) −→
g ⊙ p = p · g −1 is a left action. Which sort of action one chooses to deal with
is generally a matter of personal taste, although some actions appear more
natural in one guise than another. For example, any smooth representation
ρ : G −→ GL(V) of G on the vector space V gives rise to an action of G
on the manifold V ((g, v) −→ (ρ(g))(v)) which, because of the way matrix
multiplication is defined, it is more natural to view as acting on the left. We
will formulate the remaining definitions in terms of right actions and leave the
obvious modifications required for left actions to your imagination.
22 1. Geometrical Background
S n−1 ∼
= O(n)/O(n − 1) ∼ = SO(n)/SO(n − 1)
2n−1 ∼
S = U (n)/U (n − 1) ∼
= SU (n)/SU (n − 1)
4n−1 ∼
S = Sp(n)/Sp(n − 1).
1.3. Principal Bundles 23
where ψ : P −1 (V ) −→ G satisfies
ψ (p · g) = ψ(p)g
Remark: When n = 2 one can identify C P1 with the 2-sphere (see the
Remark on page 6) and thereby obtain a U (1)-bundle over S 2 generally
known as the complex Hopf bundle. There are, in fact, two natural ways
of identifying C P1 with S 2 and these yield bundles which are not “equiv-
alent” in a sense soon to be made precise. Since we will require both of
these bundles we shall write out their descriptions explicitly. We use the
charts (U1 , φ1 ) and (U2 , φ2 ) on C P1 (Example #5, page 5) and the stereo-
graphic projection chart (US , φS ) on S 2 , (Example #4, page 4). On its do-
main, φ−1 S ◦ φ2 is a diffeomorphism into S given by φS ◦ φ2 ([z , z ]) =
2 −1 1 2
−1 1 2 −1 −1 z 1
φS (z (z ) ) = φS ( z2 ) = (z z̄ + z̄ z , −i z z̄ + i z̄ z , |z | − |z | ). But
1 2 1 2 1 2 1 2 1 2 2 2
this last formula does not require z 2 ̸= 0 and, in fact, defines a diffeomorphism
[z 1 , z 2 ] −→ (z 1 z̄ 2 + z̄ 1 z 2 , −i z 1 z̄ 2 +i z̄ 1 z 2 , |z 1 |2 − |z 2 |2 ) of all of C P1 onto S 2 .
Composing with the projection P of S 3 onto C P1 we obtain a map
P1 : S 3 −→ S 2
given by
P1 (z 1 , z 2 ) = (z 1 z̄ 2 + z̄ 1 z 2 , −i z 1 z̄ 2 + i z̄ 1 z 2 , |z 1 |2 − |z 2 |2 )
P
and this provides a concrete realization of the bundle U (1) ,→ S 3 −→ C P1
where C P1 is identified with S 2 :
P
U (1) ,→ S 3 −→
1
S2
( 2)
Similarly, φ−1
S ◦ φ1 ([z 1 , z 2 ]) = φ−1S
z
z1 = (z̄ 1 z 2 + z 1 z̄ 2 , −i z̄ 1 z 2 + i z 1 z̄ 2 ,
|z 1 |2 − |z 2 |2 ) actually determines a global diffeomorphism of C P1 onto S 2
and thus a map
P−1 : S 3 −→ S 2
given by
P−1 (z 1 , z 2 ) = (z̄ 1 z 2 + z 1 z̄ 2 , −i z̄ 1 z 2 + i z 1 z̄ 2 , |z 1 |2 − |z 2 |2 )
Notice that P−1 (z 1 , z 2 ) differs from P1 (z 1 , z 2 ) only in the sign of the second
coordinate so P−1 is P1 followed by reflection across the “xz-plane.”
4. Let P = S 4n−1 , G = Sp(1) and σ : S 4n−1 ×Sp(1) −→ S 4n−1 the right action
σ(p, g) = p · g = (q 1 , . . . , q n ) · g = (q 1 g, . . . , q n g) described in Example #7,
page 28. The orbit space is X = H Pn−1 and we let P : S 4n−1 −→ H Pn−1 be
the quotient map. Then P(p · g) = P(p) for all p ∈ S 4n−1 and g ∈ Sp(1). For
each k = 1, . . . , n, let Vk = {[p] = [q 1 , . . . , q n ] ∈ H Pn−1 : q k ̸= 0} and define
Ψk : P −1 (Vk ) −→ Vk × Sp(1) by Ψk (p) = Ψk (q 1 , . . . , q n ) = ([p], q k /|q k |).
26 1. Geometrical Background
gji : Vi ∩ Vj −→ G
by
gji (x) = (ψj (p)) (ψi (p))−1
for any p ∈ P −1 (x). These maps are smooth and are called the transition
P
functions for G ,→ P −→ X corresponding to {(Vj , Ψj )}j∈J . They satisfy all
of the following conditions:
gii (x) = e
gij (x) = (gji (x))−1
gkj (x)gji (x) = gki (x) (cocycle condition)
For future reference we note that V1 and V2 are also the standard coordinate
neighborhoods on H P1 (called U1 and U2 in Example #5, page 5) and that
the corresponding diffeomorphisms φk : Vk −→ H = R4 are
φ2 ◦ φ−1
1 (q) = q
−1
= φ1 ◦ φ−1
2 (q)
P
for all q ∈ H − {0}. All of this is the same for U (1) ,→ S 3 −→ C P1 .
P1 P2
Let G ,→ P1 −→ X1 and G ,→ P2 −→ X2 be two principal G-bundles and,
for convenience, denote the actions of G on P1 and P2 by the same dot · . A
(principal) bundle map from P1 to P2 is a smooth map f : P1 −→ P2 which
satisfies f (p·g) = f (p)·g for all p ∈ P1 and g ∈ G. Such an f carries each fiber
of P1 diffeomorphically onto some fiber of P2 and therefore induces a smooth
map f¯ : X1 −→ X2 defined by P2 ◦f = f¯◦P1 . If X1 = X2 = X, then a bundle
map f : P1 −→ P2 is called an equivalence if it is a diffeomorphism and in-
duces the identity map on X, i.e., f¯ = idX . In this case the bundles G ,→
P1 P2
P1 −→ X and G ,→ P2 −→ X are said to be equivalent. It then follows that
−1 P
f : P2 −→ P1 is also an equivalence. If G ,→ P −→ X is a single
28 1. Geometrical Background
The Classification Theorem: Let G be a connected Lie group. Then the set
of equivalence classes of principal G-bundles over S n , n ≥ 2, is in one-to-one
correspondence with the elements of the homotopy group πn−1 (G). Thus, for
example, the principal U (1)-bundles over S 2 are in one-to-one correspondence
with the elements of π1 (U (1)) ∼ = π1 (S 1 ) ∼
= Z (the integers). Similarly, the
4
Sp(l)-bundles over S are in one-to-one correspondence with the elements of
π3 (Sp (1)) ∼
= π3 (S 3 ) ∼
= Z. We will find that the theory of characteristic classes
provides a natural means of associating an integer (Chern number) with a
bundle of either of these types.
P
For any smooth principal bundle G ,→ P −→ X the fibers P −1 (x), x ∈ X,
are all submanifolds of P diffeomorphic to G (page 260, [N4]) so, for each
p ∈ P, Tp (P ) contains a subspace isomorphic to the Lie algebra G of G
(all tangent vectors at p to smooth curves in the fiber containing p). We call
this the vertical subspace of Tp (P ) and denote it Vertp (P ). The elements of
Vertp (P ) are called vertical vectors at p. The action σ of G on P provides a
1.4. Connections and Curvature 29
natural means of identifying each Vertp (P ) with G. To see this, fix an element
A ∈ G (thought of as a set of matrices). We associate with A a vector field
A# on P , called the fundamental vector field on P determined by A, as
follows: For each p ∈ P the map σp : G −→ P defined by σp (g) = σ(p, g) = p·g
is smooth and so has a derivative (σp )∗id at the identity. Then,
d
A# (p) = (σp )∗id (A) = (p · exp(tA))|t=0
dt
(page 287, [N4]). The mapping A −→ A# (p) is an isomorphism of G onto
Vertp (P ) (Corollary 5.8.9, [N4]). Furthermore, for any A, B ∈ G,
ω p (A# (p)) = A.
Aj = adg−1 ◦ Ai + gij ∗ Θ
ij
Tp (P ) = Horp (P ) ⊕ Vertp (P )
Remark: The reason these connections are called “flat” will emerge when
we discuss the “curvature” of a connection.
But
(page 269, [N4]). A simple calculation with the coordinate formula for pullbacks
on page 9 gives 1
(sN ◦ φ−1 )∗ ω = − i(1 − cos ϕ) dθ
2
(page 270, [N4]). Similarly,
1
(sS ◦ φ−1 )∗ ω = i(1 + cos ϕ) dθ.
2
1.4. Connections and Curvature 33
P−1
Analogous calculations for U (1) ,→ S 3 −→ S 2 give results that differ from
these only by a sign.
3. (Natural connection on the quaternionic Hopf bundle) We consider the
Sp(1)-bundle
P
Sp(1) ,→ S 7 −→ H P1
(the n = 2 case of Example #4, page 31). Regard S 7 as the submanifold of H2
consisting of those (q 1 , q 2 ) with |q 1 |2 +|q 2 |2 = 1 and identify the Lie algebra of
Sp(1) with the algebra Im H of pure imaginary quaternions. Define an Im H-
valued 1-form ω̃ on H2 by ω̃ = Im (q̄ 1 dq 1 + q̄ 2 dq 2 ) and let ω be the restriction
P
of ω̃ to S 7 . Then ω is a connection form on Sp (1) ,→ S 7 −→ HP1 ((5.9.10)
and (5.9.11), [N4]). For each p ∈ S 7 , Horp (S 7 ) is that part of the real or-
thogonal complement of Vertp (S 7 ) in H2 = R8 that lies in Tp (S 7 ) (page 335,
[N4] ). The standard trivializations {(V1 , Ψ1 ), (V2 , Ψ2 )} of the bundle give
cross-sections s1 : V1 −→ P −1 (V1 ) and s2 : V2 −→ P −1 (V2 ) and we are in-
terested in the gauge potentials A1 = s∗1 ω and A2 = s∗2 ω. Since V1 and V2
are also coordinate neighborhoods for the standard charts φ1 : V1 −→ H and
φ2 : V2 −→ H we may compute these gauge potentials A1 and A2 in terms
of φ1 and φ2 coordinates. The results are as follows (pages 297–301, [N4]):
( )
q̄
(s1 ◦ φ−1 ) ∗
ω = Im dq
1
1 + |q|2
and
( )
−1 ∗ q̄
(s2 ◦ φ2 ) ω = Im dq .
1 + |q|2
To compare the two gauge potentials A1 and A2 one expresses both of them
in terms of the same coordinates on V1 ∩ V2 . For example, for any x ∈ V1 ∩ V2
and any X ∈ Tx (H P1 ),
( )
∗ φ1 (x)v̄
(s2 ω)x (X ) = Im ,
|φ1 (x)|2 (1 + |φ1 (x)|2 )
The gauge potentials A1 = s∗1 ω and A2 = s∗2 ω arose first in the physics
literature [BPST] as solutions to the Yang-Mills equations (Section 6.3, [N4]).
There they were called pseudoparticles. Today it is more common to refer
to them (or the natural connection ω from which they arose and which they
uniquely determine) as instantons.
P
4. (More instantons on Sp(1) ,→ S 7 −→ H P1 ) Let ω denote the natural
P
connection on Sp(1) ,→ S 7 −→ H P1 described in the previous example. For
P
any bundle map f : S 7 −→ S 7 , f ∗ ω is also a connection on Sp(1) ,→ S 7 −→
H P (page 37). By judiciously selecting bundle maps f (pages 336–341, [N4] )
1
one can produce, for each pair (λ, n) ∈ (0, ∞) × H, a connection form ω λ,n
P
on Sp(1) ,→ S 7 −→ H P1 , uniquely determined by the two gauge potentials
( )
( )∗ q̄ − n̄
s2 ◦ φ−1 ω λ,n = Im dq
2
λ2 + |q − n|2
and ( )
( )∗ (|n|2 + λ2 ) q̄ − n
s1 ◦ φ−1 ω λ,n = Im dq .
1
λ2 |q|2 + |1 − nq|2
P
Remark: Any connection η on Sp(1) ,→ S 7 −→ H P1 is uniquely deter-
mined by either one of the gauge potentials s1∗ η or s2∗ η (page 339, [N4]) so
it is customary to specify ω λ,n by giving the single potential
( )
q̄ − n̄
Aλ,n = Im dq .
λ2 + |q − n|2
Aλ,n is called the generic BPST potential with center n and scale λ
(page 357, [N4]).
P
The curvature Ω of a connection ω on G ,→ P −→ X is its covariant
exterior derivative, defined by having d ω operate only on horizontal parts,
i.e., for each p ∈ P and all v, w ∈ Tp (P ) we let
1.4. Connections and Curvature 35
F = dA + A ∧ A
(page 350, [N4]). Assuming (as we may) that the domain V of s is also a co-
ordinate neighborhood for a chart (V, φ) with coordinate functions x1 , . . . , xn ,
we can write A = Aα dxα and F = 12 Fαβ dxα ∧ dxβ , where the Aα and Fαβ
are G-valued functions on V . Then
Fαβ = ∂α Aβ − ∂β Aα + [Aα , Aβ ],
where we have written ∂α for ∂x∂α and these derivatives are computed
componentwise in G (Exercise 6.2.9, [N4]). If sj : Vj −→ P −1 (Vj ) and
si : Vi −→ P −1 (Vi ) are two local cross-sections with Vj ∩ Vi ̸= ∅ and if
gij : Vj ∩ Vi −→ G is the transition function relating the corresponding triv-
ializations, then sj (x) = si (x) · gij (x) for each x ∈ Vj ∩ Vi (Exercise 4.3.5,
−1 −1
[N4]) and, whereas A j = gij A i gij + gij dgij , we have
−1
F j = gij F i gij .
Ω = dω.
−1
If s : V −→ P (V ) is any cross-section, then we can write the gauge potential
A = s∗ ω and field strength F = s∗ Ω as
A = −i A
F = d A = −i dA = −i F ,
where A and F are real-valued forms on V (the minus signs are conventional).
If sj : Vj −→ P −1 (Vj ) and si : Vi −→ P −1 (Vi ) are two local cross-sections
with Vj ∩ Vi ̸= ∅ and if gij : Vj ∩ Vi −→ U (1) is the corresponding transition
function, then
−1 −1 −1
Aj = gij Ai gij + gij dgij = Ai + gij dgij
and
−1
F j = gij F i gij = F i
on Vj ∩ Vi because U (1) is Abelian. In particular, the local field strengths,
since they agree on any intersections of their domains, piece together to give
a globally defined field strength 2-form F on X. This is a peculiarity of Abelian
gauge fields and generally is not true in the non-Abelian case. Also note that
since gij maps into U (1) it can be written as gij (x) = e−i Λ(x) for some real-
−1
valued function Λ on Vj ∩ Vi . Then gij dgij = ei Λ e−i Λ (−i dΛ) = −i dΛ so
Aj = Ai − i dΛ, i.e.,
Aj = Ai + dΛ,
which is the traditional form for the relationship between two “vector
potentials.”
1.4. Connections and Curvature 37
(see the Remark on page 42). A calculation (pages 327–328, [N4]) gives
1 λ2
F λ,n = d Aλ,n + [Aλ,n , Aλ,n ] = 2 d q̄ ∧ dq
2 (λ + |q − n|2 )2
YM (Aλ,n ) = 8π 2 .
P
Remarks: All of the connections ω λ,n on Sp(1) ,→ S 7 −→ HP1 give rise to
potentials Aλ,n with the same Yang-Mills action YM(Aλ,n ) and this, we will
find, is no accident. The number 8π1 2 YM(Aλ,n ) is, in fact, a topological char-
acteristic of the quaternionic Hopf bundle, not unlike the Euler characteristic
of a surface. We will eventually show that this number is essentially the inte-
gral over HP1 of what is called the 2nd Chern class of the bundle. Like the
magnetic charge of a Dirac monopole, the Yang-Mills action for an instanton
can be thought of as a sort of “topological charge.”
P
is any local trivialization of G ,→ P −→ X and s : V −→ P −1 (V ) is the
associated cross-section, then the map Φ̃ : V × F −→ PG−1 (V ) defined by
−1
Φ̃(x, ξ) = [s(x), ξ] is a homeomorphism with inverse Ψ̃ : PG (V ) −→ V × F
given by Ψ̃([s(x), ξ]) = (x, ξ) (page 381, [N4]). If (Vi , Ψi ) and (Vj , Ψj ) are two
such trivializations with Vi ∩Vj ̸= ∅ and gji : Vi ∩Vj −→ G is the corresponding
transition function, then Ψ̃j ◦ Ψ̃−1 i : (Vi ∩ Vj ) × F −→ (Vi ∩ Vj ) × F is given by
Ψ̃j ◦ Ψ̃−1
i (x, ξ) = (x, g ji (x) · ξ) and so is a diffeomorphism (page 382, [N4]).
It follows that there is a unique differentiable structure on P ×G F relative to
−1
which each Ψ̃ : PG (V ) −→ V × F is a diffeomorphism and that, relative to
this structure, PG : P ×G F −→ X is smooth. We call
PG : P ×G F −→ X
P
the fiber bundle associated with G ,→ P −→ X by the given left action
of G on F .
The special case of most interest to us arises as follows: Let F = V be a
finite dimensional vector space (with its natural differentiable structure) and
ρ : G −→ GL(V) a smooth representation of G on V. Then ρ gives rise to a
smooth left action of G on V ((g, v) −→ g · v = (ρ(g))(v)). The fiber bundle
P
associated with G ,→ P −→ X by this action is denoted
Pρ : P ×ρ V −→ X
P
and called the vector bundle associated with G ,→ P −→ X by
the representation ρ. In this case each fiber Pρ−1 (x) = {[p, v] : v ∈ V},
where p is any point in P −1 (x), is a copy of V and admits a natural vector
space structure: a1 [p, v1 ] + a2 [p, v2 ] = [p, a1 v1 + a2 v2 ] for all a1 , a2 ∈ R and
v1 , v2 ∈ V. We record a few examples of particular interest.
P
1. Let U (1) ,→ P −→ X be an arbitrary principal U (1)-bundle and
take V = C (as a 2-dimensional real vector space). If ρ : U (1) −→ GL(C)
is any representation of U (1) on C, then the associated vector bundle
Pρ : P ×ρ C −→ X has fibers that are copies of C and is called a
complex line bundle over X. An obvious choice for ρ : U (1) −→ GL(C)
is obtained by taking (ρ(g))(z) = gz for each g ∈ U (1) and z ∈ C (if
g = ei θ , 0 ≤ θ < 2π, then ρ(g) is rotation by θ). More generally, one can define,
for any integer n, a representation ρ : U (1) −→ GL(C) by (ρ(g))(z) = g n z
and thereby an associated vector bundle over X.
P
2. Let Sp(1) ,→ P −→ X be any principal Sp(1)-bundle and take V = H
(as a 4-dimensional real vector space). Then any representation ρ : Sp(1) −→
GL(H) (e.g., (ρ(g))(q) = gq, or (ρ(g))(q) = g n q for some integer n) defines a
vector bundle P ×ρ H with fibers isomorphic to H and called a quaternionic
line bundle over X.
40 1. Geometrical Background
P
3. Let SU (2) ,→ P −→ X be any principal SU (2)-bundle and take
V = C2 (as a 4-dimensional real vector space). Then any representation
ρ : SU (2) −→ GL(C2 ) defines an associated vector bundle P ×ρ C2 . A par-
ticularly useful choice (for )ρ is obtained as (
follows
) (we write the elements of
C as column vectors ξ2 ): For each g = γ δ ∈ SU (2) define
2 ξ1 αβ
( ) ( ) ( ) ( )
ξ1 ξ1 α β ξ1 αξ 1 + βξ 2
(ρ(g)) =g = = .
ξ2 ξ2 γ δ ξ2 γξ 1 + δξ 2
1
This representation of SU (2) is generally denoted D 2 and we will find (in
Section 2.4) that it arises naturally in Pauli’s nonrelativistic theory of the
electron.
P
4. Let G be an arbitrary matrix Lie group and G ,→ P −→ X be an ar-
bitrary principal G-bundle. The adjoint representation ad : G −→ GL(G)
assigns to each g ∈ G the nonsingular linear transformation adg on the
Lie algebra G defined by
adg (A) = g A g −1
P
(page 24). The vector bundle associated with G ,→ P −→ X by ad is called
P
the adjoint bundle of G ,→ P −→ X and denoted
ad P = P ×ad G.
P
Now, let G ,→ P −→ X be an arbitrary principal G-bundle, F a smooth
manifold on which G acts on the left and PG : P ×G F −→ X the as-
sociated fiber bundle. If V is an open subset of X, then a smooth map
ϕ : P −1 (V ) −→ F is said to be equivariant (with respect to the given actions
of G on P and F ) if
ϕ(p · g) = g −1 · ϕ(p)
x −→ [s(x), ψ(x)].
42 1. Geometrical Background
Then, in gauge s · g,
σg∗ (dω ϕ) = g −1 · dω ϕ
for each g ∈ G ((6.8.1), [N4]). These are the derivatives that appear in the
field equations describing the quantitative response of the particle to the gauge
field (pages 391–392, [N4]). A computational formula analogous to the Cartan
Structure Equation for curvature Ω (which is the covariant exterior derivative
of the connection form ω) is obtained as follows: For any A ∈ G and v ∈ V we
define A · v ∈ V by
d d
A·v = (exp(tA) · v)|t=0 = (ρ(exp(tA))(v))|t=0 .
dt dt
Remarks: Two special cases are worth pointing out immediately. If G is
a group of n × n matrices (with entries in F = R, C, or H), V = Fn
(thought of as column matrices) and ρ is the natural representation of G
on V (matrix multiplication), then, identifying G with an algebra of matrices,
A · v = Av (matrix multiplication). On the other hand, if V = G and ρ = ad,
then, for all A, B ∈ G, A · B = [A, B] (see the proof of Theorem 5.8.8,
[N4]).
1.5. Associated Bundles and Matter Fields 43
dω ϕ = dϕ + ω · ϕ
ρ : G −→ GL(V)
ψ (p · g) = g −1 · ψ (p)
U : V −→ R
2.1. General Framework for Classical Gauge Theories 47
U (g · v) = U (v).
We will spell out in detail what each of these terms means in the concrete ex-
amples to follow. Briefly, c is a normalizing constant, c 1 and c 2 are “coupling
constants,” Fω is a global 2-form on X with values in the ad-joint bundle ad P
which locally pulls back to the gauge field strengths F , dω ϕ is the covariant
exterior derivative of the matter field ϕ (thought of as a cross-section of the
associated vector bundle) and the norms arise from the metric on X and the
Killing form on the Lie algebra G of G. Integrals of such objects over mani-
folds like X will be introduced in Chapter 4. The physically interesting field
configurations (ω, ϕ), are those which (at least locally) minimize the value
of the action functional. The Calculus of Variations provides necessary con-
ditions (the Euler-Lagrange differential equations) that must be satisfied by
such minima. The Euler-Lagrange equations for the action A(ω, ϕ), are the
appropriate field equations (the “equations of motion”) of our gauge theory.
One can, of course, generalize the model we have described by including more
than one matter field.
From the point-of-view of physics, one is generally interested only in finite
action configurations (ω, ϕ), i.e., those for which
A (ω, ϕ) < ∞.
This is assured if X is compact, but otherwise one must assume some sort of
appropriate asymptotic behavior for the terms in the integrand. Such asymp-
totic conditions turn out to have profound topological consequences (e.g., the
existence of “topological charge”) and investigating this link between topology
and asymptotics is one of our primary objectives.
These then are the basic ingredients required to build a classical gauge
theory. There is a special case that has gotten a great deal of attention, par-
ticularly in the mathematical community. When c1 = c2 = 0 in the action
A (ω, ϕ), (so that, effectively, there are no matter fields present), then one can
think of A as depending only on ω and, in this case, it is referred to as the
48 2. Physical Motivation
d ω Fω = 0.
These last two equations lie at the heart of what is called pure Yang-Mills
theory. The impact of this subject on low dimensional topology is discussed
at some length in [N4]. Although this special case may not appear to be in
the spirit of our announced intention here to model interactions we will find
that it leads (through a process known as “dimensional reduction”) directly to
the particular interactions of most interest to us in Section 2.5.
Remark: For the moment we will require very little of the geometry of R1,3
and its physical significance. Simply think of the elements of R1,3 as “events”
whose standard coordinates represent the time (x0 ) and spatial (x1 , x2 , x3 ) co-
ordinates by which the event is identified in some fixed, but arbitrary inertial
frame of reference. The entire history of a (point) object can then be identified
with a continuous sequence of events (i.e., a curve) in R1,3 called its “world-
line.” Finally, since the differentiable structure of R4 is just its natural struc-
ture as a real vector space (Example #3, page 4), each tangent space Tp (R1,3 )
is canonically identified with R4 itself. Since the components of the semi-
Riemannian metric we have introduced are the same at every p ∈ R1,3 one
can think of R1,3 simply as the vector space R4 equipped with the Minkowski
inner product g(v, w) = ηαβ v α wβ = v 0 w0 − v 1 w1 − v 2 w2 − v 3 w3 . A vector
v in R1,3 is said to be spacelike, timelike, or null if g(v, v) is < 0, > 0,
or = 0, respectively. The physical origin of the terminology will emerge as we
procced.
We introduce a matrix
1 0 0 0
0 −1 0 0
η = (ηαβ ) =
0 0 −1 0
0 0 0 −1
over X and a connection ω on it (we consider first the pure Yang-Mills theory
in which matter fields are absent). Since U (1) is Abelian, all brackets in the Lie
algebra u(1) = Im C are zero so the curvature Ω of ω is given by Ω = dω.
If s : V −→ P is a local cross-section, then we may write the local gauge
potential and field strength as
A = s∗ ω = −i A
F = s∗ Ω = d A = −i d A = −i F ,
and
F j = gij −1 F i gij = F i
on Vj ∩ Vi because U (1) is Abelian. In particular, the local field strengths,
since they agree on any intersections of their domains, piece together to give
a globally defined field strength 2-form F on X. This is a peculiarity of Abelian
gauge theories and one should note that, even here, the potentials A do not
agree on the intersections of their domains and so do not give rise to a globally
defined object on X. Indeed, since the transition function gij is a map into
U (1) it can be written as
gij (x) = e−i Λij (x)
so that gij −1 dgij = −i dΛij and Aj = Ai − i dΛij . Equivalently,
Aj = Ai + dΛij ,
which is the traditional form for the relationship between two “vector potentials.”
Relative to standard coordinates x0 , x1 , x2 , x3 on R1,3 we can write, for any
s : V −→ P ,
A = Aα dxα = −i Aα dxα
and
1 1
F = Fαβ dxα ∧ dxβ = − i Fαβ dxα ∧ dxβ ,
2 2
where
Fαβ = ∂α Aβ − ∂β Aα + [Aα , Aβ ]
= ∂α Aβ − ∂β Aα (because U (1) is Abelian)
= −i (∂α Aβ − ∂β Aα )
= −i Fαβ .
The Fαβ are skew-symmetric in α and β. To make some contact with the
notation used in physics we define functions E 1 , E 2 , E 3 and B 1 , B 2 , B 3 by
Fi0 = Ei
and
Fij = εijk B k
where i, j, k = 1, 2, 3 and εijk is the Levi-Civita symbol (1 if ijk is an even
permutation of 123, −1 if ijk is an odd permutation of 123, and 0 otherwise).
Thus,
0 −E 1 −E 2 −E 3
1
E 0 B 3 −B 2
(Fαβ ) = 2
E −B 3 0 B1
E 3
B 2
−B 1
0
and
2.2. Electromagnetic Fields 51
1
F = Fαβ dxα ∧ dxβ = −E 1 dx0 ∧ dx1 − E 2 dx0 ∧ dx2 − E 3 dx0 ∧ dx3
2
+ B 3 dx1 ∧ dx2 − B 2 dx1 ∧ dx3 + B 1 dx2 ∧ dx3
= (E 1 dx1 + E 2 dx2 + E 3 dx3 ) ∧ dx0
+ B 3 dx1 ∧ dx2 + B 1 dx2 ∧ dx3 + B 2 dx3 ∧ dx1 .
⃗ = (E 1 , E 2 , E 3 ) and B
One is to think of E ⃗ = (B 1 , B 2 , B 3 ) as the “electric field”
and the “magnetic field,” respectively, that correspond to F (the justification
for thinking this way will appear shortly).
Next we introduce functions F αβ on X defined by
F αβ = η αγ η βδ Fγδ , α, β = 0, 1, 2, 3
∗ 1
Fαβ = εαβγδ F γδ , α, β = 0, 1, 2, 3.
2
Writing these out one finds that
0 B1 B2 B3
−B 1 0 E3 −E 2
(∗ Fαβ ) =
−B 2
−E 3 0 E1
−B 3 E2 −E 1 0
so that
∗ 1 ∗
F = F αβ dxα ∧ dxβ
2
= (−B 1 dx1 − B 2 dx2 − B 3 dx3 ) ∧ dx0
+ E 3 dx1 ∧ dx2 + E 1 dx2 ∧ dx3 + E 2 dx3 ∧ dx1 .
We also define
∗
F = −i ∗F .
Finally, one can also “raise the indices” of ∗F and define ∗F αβ = η αγ η βδ ∗Fαβ
so that
0 −B 1 −B 2 −B 3
1
B 0 E 3 −E 2
∗
( Fαβ ) = 2 .
B −E 3 0 E1
B 3
E 2
−E 1
0
One can think of the Hodge dual as the 2-form obtained by formally replacing
B⃗ by E⃗ and E ⃗ by −B.
⃗
All of this apparently ad hoc notation will eventually be seen to fit naturally
into the general scheme of things. For the time being the reader may wish to
regard all of it as simply a useful bookkeeping device. For example, one has
the following easily verified formulas:
1
Fαβ F αβ = |B|⃗ 2 − |E|
⃗ 2
2
1
Fαβ ∗ F αβ = E
⃗ ·B⃗
4
for the two scalar invariants normally associated with F in classical electro-
magnetic theory.
But what is the justification for all of these references to classical elec-
tromagnetic theory? We began by looking at an arbitrary connection ω on
an arbitrary principal U (1)-bundle over an open submanifold of Minkowski
spacetime and have deviously interjected things we have called “electric fields”
and “magnetic fields.” Is there any reason to believe that these objects have
anything whatever to do with what physicists call “electric fields” and “mag-
netic fields?” The answer lies in the Yang-Mills equations. We are, after all,
not really interested in arbitrary connections, but only in the stationary val-
ues of the Yang-Mills action (page 56). In our present circumstances we will
find that this action can be written
∫
1
YM(ω) = − Fαβ F αβ dx0 dx1 dx2 dx3
X 4
and the corresponding Yang-Mills equations are
d ∗F = 0,
while the Bianchi identity is
dF = 0
(in the Abelian case, covariant exterior derivatives are just ordinary exterior
derivatives). In standard coordinates these read
2.2. Electromagnetic Fields 53
∂α F αβ = 0, β = 0, 1, 2, 3
and
∂α ∗ F αβ = 0, β = 0, 1, 2, 3,
respectively. Now, the remarkable part is that, if one writes these out in terms
⃗ and B’s
of the E’s ⃗ we introduced earlier the result is
⃗
∇ ⃗ − ∂ E = ⃗0
⃗ ×B and ∇
⃗ ·E
⃗ =0 (d ∗F = 0)
∂x0
and
⃗
∇ ⃗ + ∂ B = ⃗0
⃗ ×E and ∇
⃗ ·B
⃗ =0 (dF = 0),
∂x0
Now, the remarkable part of all this is that this cohomology class (not the
2-form, but its cohomology class) does not depend on the initial choice of
the connection ω from which it arose. It is therefore a characteristic of the
bundle itself and not of the connection. Indeed, c1 (P ) is the simplest example
of what is called a characteristic class for the bundle. We’ll encounter one
more example of such a thing (the second Chern class) in Section 2.5.
Now let us specialize to the case of U (1)-bundles. Here we have seen that
there is a globally defined field strength F for any connection. Moreover, since
u(1) consists of 1 × 1 matrices, the trace just picks out the sole entry in this
matrix so trace F = −iF and therefore
2.2. Electromagnetic Fields 55
[ ]
i 1
c1 (P ) = (−iF ) = [F ] (G = U (1)).
2π 2π
In particular, the first Chern class for a principal U (1)-bundle over an open
1
submanifold of Minkowski spacetime is just 2π times the cohomology class
of the globally defined (electromagnetic) field F on X. Since principal U (1)-
bundles over any manifold are classified up to equivalence by their first Chern
classes (Appendix E, [FU]), we conclude that the U (1)-bundle on which an
electromagnetic field F is to be modeled as a connection is uniquely deter-
mined by the cohomology class of F .
Now we return to our concrete examples. First, the Coulomb field, i.e., a
static, purely electric field of a point charge which we assume to be located at
the (x1 , x2 , x3 )-origin in R1,3 . Thus, the worldline of our source is the x0 -axis
in R1,3 so we take
X = R1,3 − {(x0 , 0, 0, 0) ∈ R1,3 : x0 ∈ R}.
Define A = Aα dxα = (−n/ρ)dx0 , where n is an integer and ρ > 0 with
ρ2 = (x1 )2 + (x2 )2 + (x3 )2 (we measure “charge” in multiples of the charge of
the electron so it is an integer). A simple calculation shows that the functions
Fαβ = ∂α Aβ − ∂β Aα , α, β = 0, 1, 2, 3, are given by
0 −x1 −x2 −x3
n
x1 0 0 0
(Fαβ ) = 3 2 .
ρ x 0 0 0
x3 0 0 0
Thus,
n
F = 3 (x1 dx1 + x2 dx2 + x3 dx3 ) ∧ dx0
ρ
so
B⃗ = ⃗0 and E ⃗ = n ⃗r, ⃗r = (x1 , x2 , x3 ).
ρ3
This is, of course, the classical Coulomb field that we wish to describe.
The critical observation here is this: Our Coulomb potential
A = (−n/ρ)dx0 is defined and satisfies dA = F globally on all of X so
that F is exact on X and its cohomology class [F ] ∈ HdeR 2
(X) is zero. Thus,
the U(1)-bundle on which F is modeled by a connection with field strength
F = −iF has first Chern class zero and so must be the trivial bundle. This is
true for any charge n so that, in particular, the electric charge of the source
is not encoded in the topology of this bundle (we will find that the situation
is quite different for a Dirac magnetic monopole).
Here’s another way to look at this: Somewhat later we will calculate the
cohomology of X = R1,3 − {(x0 , 0, 0, 0) ∈ R1,3 : x0 ∈ R} and find that
{
k
R, k = 0, 2
HdeR (X) = .
0, otherwise
56 2. Physical Motivation
0 x2 −x1 0
Thus, ∗ n
F = 3 (x1 dx2 ∧ dx3 − x2 dx1 ∧ dx3 + x3 dx1 ∧ dx2 ).
ρ
Notice that, on the 2-sphere ρ = 1 in any x0 = constant slice of X, ∗ F reduces
to n times x1 dx2 ∧ dx3 − x2 dx1 ∧ dx3 + x3 dx1 ∧ dx2 on S 2 . This, as we shall
see, is what is called the standard volume form of S 2 and its integral over S 2
is the area 4π of S 2 (Section 4.6). Thus, the integral of ∗ F over this sphere
is 4πn, which is nonzero and this implies that ∗ F cannot be cohomologically
trivial. Thus, [∗ F ] generates HdeR2
(X). Furthermore, we will show that the
integral of F over any 2-sphere surrounding {(x0 , 0, 0, 0) : x0 ∈ R} is the
∗
same so these integrals “detect” the charge n enclosed by the sphere. Over
any 2-sphere that does not enclose the x0 -axis, the integral of ∗ F is zero (this
will follow from “Stokes’ Theorem”).
The electric charge n is not “topological” because all Coulomb fields are
represented by connections on the trivial bundle—the charge is not encoded
in the topology of the bundle. The situation is quite different for a Dirac
monopole (which is not surprising since the Hodge dual for 2-forms on X
essentially interchanges “electric” and “magnetic” so, for a magnetic charge,
[F ] will play the role that [∗ F ] played for the Coulomb field). In more detail,
we once again let X = R1,3 − {(x0 , 0, 0, 0) ∈ R1,3 : x0 ∈ R}. For the moment
we let g denote an arbitrary real number (to be thought of as the magnetic
“charge” of the monopole whose worldline is the x0 -axis). We are interested
in the field F = 12 Fαβ dxα ∧ dxβ , where
0 0 0 0
g
0 0 x3
−x2
(Fαβ ) = 3 .
ρ 0 −x3 0 x1
0 x2 −x1 0
Thus,
g
F = (x1 dx2 ∧ dx3 − x2 dx1 ∧ dx3 + x3 dx1 ∧ dx2 )
ρ3
2.2. Electromagnetic Fields 57
so
E ⃗ = g ⃗r,
⃗ = ⃗0 and B ⃗r = (x1 , x2 , x3 ).
ρ3
Since F is independent of x0 this defines a 2-form on any x0 = constant
slice. Moreover, expressed in terms of standard spherical coordinates (ρ, φ, θ)
on such a slice,
F = g sin φ dφ ∧ d θ.
Notice that this is independent of ρ and so may be further restricted to the
copy ρ = 1 of S 2 in, say, the x0 = 0 slice of X. Now, if the monopole field on
X is the field strength of some connection on a U(1)-bundle over X, then its
restrictions would likewise be field strengths for connections on U(1)-bundles
over the submanifolds (restriction means pullback by the inclusion map and
the inclusion of a restricted bundle is a bundle map). Henceforth, we will
concentrate on these restrictions.
For the field F under consideration there is no globally defined potential
A satisfying dA = F (pages 2–3 of [N4]), but there are the usual local
potentials. Specifically, we define AN and AS on UN = S 2 − {(0, 0, 0, −1)}
and US = S 2 − {(0, 0, 0, 1)}, respectively, by
AN = g(1 − cos φ) dθ
and
AS = −g(1 + cos φ) dθ.
Then, on their respective domains, these satisfy dAN = F and dAS = F . Con-
sider now the corresponding u(1)-valued forms (we identify u(1) with Im C):
AN = −i AN = −i g(1 − cos φ) dθ
AS = −i AS = i g (1 + cos φ) dθ
F = −i F = −i g sin φ d φ ∧ d θ
(we assume S 2 has its standard orientation). Thus, 2g ∈ Z. From yet another
perspective, the restriction of gSN (φ, θ) = e−2gθi to the equatorial circle S 1
in UN ∩ US (i.e., ei θ −→ (ei θ )−2g ) would be the “characteristic map” whose
homotopy type determines the bundle (page 228, [N4]) and this map is not
even well-defined (single-valued) on S 1 unless 2g is an integer. However you
choose to view the situation, we will henceforth restrict our attention to the
following forms:
1
AN = − ni (1 − cos φ) dθ
2
1
AS = ni (1 + cos φ)dθ
2
1
F = − ni sin φ dφ ∧ dθ.
2
For each fixed integer n, the potentials AN and AS uniquely determine a
connection ω n on the principal U (1)-bundle
P
U (1) ,→ Pn −→
n
S2
n=1
ω 1 is given by
ω 1 = i ι∗ (Im(z̄ 1 dz 1 + z̄ 2 dz 2 )),
where ι : S 3 ,→ C2 is the inclusion.
n = −1
This gives the natural connection on the alternate version of the complex Hopf
bundle
P−1
U (1) ,→ S 3 −→ S 2 ,
where P−1 is the restriction to S 3 ⊆ C2 of the map
( ) ( 2 2
)
P−1 z 1 , z 2 = z 1 z̄ 2 + z̄ 1 z 2 , i z 1 z̄ 2 − i z̄ 1 z 2 , z 1 − z2 .
ω −1 is given by
( ( ))
ω −1 = i ι∗ Im z̄ 1 dz 1 + z̄ 2 dz 2 = ω1 ,
n=0
ω 0 = π ∗ Θ,
where π : S 2 × U (1) −→ U (1) is the projection onto the second factor and Θ
is the Cartan 1-form on U (1).
n>1
P
Denote by U (1) ,→ Pn −→ n
S 2 the principal U (1)-bundle over S 2 with transi-
−nθi
tion function gSN (φ, θ) = e . We can identify Pn explicitly as a manifold as
follows: Identify the discrete group Zn of integers modulo n with the following
subgroup of U (1):
Zn = {e2kπi /n : k = 0, 1, . . . , n − 1}.
Then Zn acts on S 3 on the right (because U (1) does). We let S 3 /Zn be the
orbit space, e.g., S 3 /Z2 = RP3 . One can provide S 3 /Zn with a manifold
60 2. Physical Motivation
structure in the same way as for RP3 . The Hopf map P1 : S 3 −→ S 2 carries
each orbit of the usual U (1)-action on S 3 to a point in S 2 so it does the same
for a Zn -orbit. Moreover, each point of S 2 is the image under P1 of a Zn -orbit
(indeed, of many Zn -orbits). Thus, P1 descends to a surjective map
Pn : S 3 / Zn −→ S 2 .
Also note that the usual U (1)-action on S 3 carries any Zn -orbit onto another
Zn -orbit which is inside the same U (1)-orbit (and so has the same image under
Pn ). Thus, the U (1)-action on S 3 descends to a U (1)- action on S 3 /Zn which
preserves the fibers of Pn . Local triviality of Pn : S 3 /Zn −→ S 2 follows so
that
Pn
U (1) ,→ S 3 / Zn −→ S2
is a principal U (1)-bundle and the transition function gSN is given
by gSN (φ, θ) = e−n θ i . Since a bundle is determined by its transition functions,
Pn
we have an explicit model for U (1) ,→ Pn −→ S 2 . In particular, Pn ∼
= S 3 / Zn .
Remark: Pn = S 3 / Z n is an example of a “lens space.”
P
Since the transition function for U (1) ,→ S 3 / Z n −→ n
S 2 is
−n θ i
gSN (φ, θ) = e , AN = − 2 ni (1 − cos φ) dθ and AS = 2 ni (1 + cos φ) dθ
1 1
ρn (g)(z) = g · z = g n z,
ϕ(p · g) = g −n ϕ(p)
∂ α = η αβ ∂β , α = 0, 1, 2, 3
Aα = η αβ Aβ , α = 0, 1, 2, 3.
Although the expression we have given for ∥dω ϕ∥2 is local and appears to
depend on the choice of the cross-section s, it is, in fact, gauge invariant and
therefore determines a globally defined, real-valued function on X. To see
2.3. Spin Zero Electrodynamics 63
this we observe the following: We have already seen (page 59) that a gauge
transformation g can be written g(x) = e−i Λ(x) and has the following effects
on the potential A and the matter field ϕ:
A −→ Ag = A + dΛ
ϕ −→ ϕg = g −1 · ϕ = ei nΛ ϕ.
Thus,
and, similarly,
Consequently,
( )
(∂α ϕg − i n(Ag )α ϕg ) ∂ α ϕg + i n(Ag )α ϕg
( )
= (∂α ϕ − i nAα ϕ) ∂ α ϕ̄ + i nAα ϕ̄
as required. With this and an appropriate choice of normalizing and coupling
constants, we take our action functional to be
∫ [
1 1
A(ω, ϕ) = − Fαβ F αβ + (∂α ϕ − i nAα ϕ) (∂ α ϕ̄ + i nAα ϕ̄)
X 4 2
]
1
+ mϕϕ̄ dx0 dx1 dx2 dx3 .
2
(see [Bl]). The first of these is the Klein-Gordon equation (for a spin zero
particle of mass m and charge n interacting with a gauge field F = −i F =
−i dA determined by the local gauge potentials A = −i Aα dxα ). The second
equation is equivalent to d∗ F = 0 (see page 62). Since −i F is the pullback
of a curvature form, the Bianchi identity gives dF = 0 also so F satisfies
Maxwell’s equations.
Remark: Of course, it was our stated intention to model a spin zero particle
in an electromagnetic field, but the point here is that the electromagnetic
64 2. Physical Motivation
nature of the field F that appears in the action A(ω, ϕ) is necessitated by the
Euler-Lagrange equations. We do not have to impose Maxwell’s equations “by
hand.” Also note that conjugating the Klein-Gordon equation gives
∂α ∂ α ϕ + m2 ϕ = 0.
where (V, A)⃗ = (A0 , A1 , A2 , A3 ) = (A0 , −A1 , −A2 , −A3 ) and A = Aα dxα
satisfies dA = F (see page 63).
Remark: The meaning of (−i ∇
⃗ − nA)
⃗ 2 as an operator on ψ is as follows:
( )2 ( ) ( )
−i ∇
⃗ − nA
⃗ ψ = −i ∇ ⃗ − nA⃗ · −i ∇⃗ − nA⃗ ψ
( ) ( )
= −i ∇⃗ − nA⃗ · −i ∇ψ
⃗ − n ψA ⃗
= −∇
⃗ 2 ψ + ni ∇⃗ · (ψ A) ⃗ · (∇ψ)
⃗ + ni A ⃗
+ n2 | A
⃗ |2 ψ.
Now, ψ takes values in C and so, at each point, has a modulus and a phase
(ψ = rei θ ). There is some arbitrariness in the phase, however, since, if a is an
element of U (1) (identified with a complex number of modulus 1), then aψ
satisfies (2.3.1) whenever ψ does (being a constant, a just slips outside of all
the derivatives in (2.3.1)). Moreover, aψ differs from ψ only in the phase fac-
tor (since |a| = 1) and so |aψ|2 = |ψ|2 . Since all of the physically significant
probabilities in quantum mechanics depend only on this squared modulus,
ψ and aψ should represent the same physical object. This freedom to alter
the phase of ψ is quite restricted, however. Since a must be constant, any
phase shift in the wavefunction must be accomplished at all spatial locations
simultaneously. Such a global phase shift, however, violates both the spirit
and the letter of relativistic law (you can’t do anything “at all spatial loca-
tions simultaneously”). Nevertheless, it is difficult to shake the feeling that
the physical significance of ψ “should” persist under some sort of phase shift
(again, because squared moduli will be unaffected). Notice that the relativistic
objection to a phase shift would disappear if one allowed the phase to shift
independently at each spacetime point, i.e., if one replaced ψ by aψ, where a
is now a function of (x, y, z, t) taking values in U (1). The problem with this,
of course, is that, if a is not constant, it will not simply “slip outside” of all
the derivatives in (2.3.1). Indeed, product rules will generate all sorts of new
terms that do not cancel so there is no reason to suppose that aψ will even
be a solution to (2.3.1). A bit (actually, quite a bit) of vector calculus will, in
fact, establish the following: Let ψ be a solution to (2.3.1) and let Λ(x, y, z, t)
be any smooth, real-valued function. Then
ψ ′ = ei n Λ ψ
66 2. Physical Motivation
is a solution to
( ( ))
1 ( ⃗ ( ))2 ∂ ∂Λ
−i ∇ − n A⃗ + ∇Λ
⃗ ψ′ = i −n V − ψ′ . (2.3.2)
2m ∂t ∂t
1 ⃗ 2 ψ = i ∂ ψ.
(−i ∇) (2.3.3)
2m ∂t
2.3. Spin Zero Electrodynamics 67
On the other hand, the same (trivial) electromagnetic field is described by any
potential of the form (0 − ∂Λ
∂t , 0 + ∇Λ). Begging the indulgence of the reader
⃗ ⃗
we would now like to call this (V, A)⃗ and write (2.3.2) as
( )
1 ( )2
⃗ ψ ′ = i ∂ − nV ψ ′ .
−i ∇
⃗ − nA (2.3.4)
2m ∂t
and ( )
∂
(∂ , ∂ , ∂ , ∂ ) = (∂0 , −∂1 , −∂2 , −∂3 ) =
0 1 2 3
, −∇
⃗
∂t
so
( ) ( )
∂ ⃗ = i ∂ + i n V, −∇
i − n V, −i ∇
⃗ − nA ⃗ + i nA⃗
∂t ∂t
( 0
= i ∂ + i n V, ∂ 1 + i n A1 ,
)
∂ 2 + i n A 2 , ∂ 3 + i n A3 .
Since ( ) ( )
∂ ∂ ( )
i , −i ∇ = i
⃗ , −∇ = i ∂ 0 , ∂ 1 , ∂ 2 , ∂ 3 ,
⃗
∂t ∂t
the substitutions above simply amount to
∂ α −→ ∂ α + i n Aα , α = 0, 1, 2, 3,
∂α −→ ∂α + i n Aα , α = 0, 1, 2, 3
(the sign is + rather than − because of our decision to include the conventional
minus sign in A = −i A).
We will conclude our discussion of the Klein-Gordon equation by de-
scribing its two most important invariance properties: gauge invariance and
Lorentz invariance. We have already shown that the action A(ω, ϕ) is invariant
under gauge transformations and it follows that the same is true of its Euler-
Lagrange equations. Nevertheless, a direct proof is instructive. Thus, we con-
sider the equation
(∂α − i n Aα ) (∂ α − i n Aα )ϕ + m2 ϕ = 0 (2.3.5)
A −→ Ag = A + d Λ
and
ϕ −→ ϕg = g −1 · ϕ = ei n Λ ϕ.
Our objective is to show that, if ϕ satisfies (2.3.5), then
First, we compute
2.3. Spin Zero Electrodynamics 69
( )
(∂ α − i n (Ag )α ) ϕg = (∂ α − i n (Aα + ∂ α Λ)) ei n Λ ϕ
= ∂ α (ei nΛ ϕ) − i n Aα ei n Λ ϕ − i n ∂ α Λ ei n Λ ϕ
= ei n Λ ∂ α ϕ + i n e i n Λ ∂ α Λ ϕ
− i n A α ei n Λ ϕ − i n ∂ α Λ ei n Λ ϕ
= ei n Λ (∂ α ϕ − i n Aα ϕ)
= ei n Λ (∂ α − i n Aα ) ϕ.
so
= ei n Λ (∂α − i n Aα )(∂ α − i n Aα ) ϕ.
From this it is clear that (2.3.5) implies (2.3.6). The Klein-Gordon equation
is gauge invariant.
We wish to show next that the Klein-Gordon equation is “relativistically
invariant.” Roughly, this means that the equation has the same mathematical
form in all inertial frames of reference, but the precise meaning of such a
statement in general will require some discussion. In this section we will be
content to spell out explicitly what is being asserted for the Klein-Gordon
equation alone. When we turn to the Dirac equation in the next section we
will describe precisely what is meant by “relativistic invariance” in general.
Thus far we have written the Klein-Gordon equation (2.3.5) only in
standard coordinates x0 , x1 , x2 , x3 for R1,3 , i.e., only in one fixed inertial frame
of reference. To emphasize this fact we will now write the derivatives ∂α and
∂ α explicity as ∂/∂xα and η αβ ∂/∂xβ so that (2.3.5) becomes
( )(
∂ ∂
η αβ α
− i n A α (x 0
, . . . , x3
) − i n Aβ (x0 ,
∂x ∂xβ
) (2.3.7)
. . . , x3 ) ϕ (x0 , . . . , x3 ) + m2 ϕ (x0 , . . . , x3 ) = 0.
The basic postulate of Special Relativity is that one inertial frame of reference
is as good as another. The coordinates y 0 , y 1 , y 2 , y 3 for R1,3 , supplied by
another such frame of reference are assumed to be related to x0 , x1 , x2 , x3 by
y α = Λα β x β , α = 0, 1, 2, 3,
70 2. Physical Motivation
x β = Λα β y α , β = 0, 1, 2, 3.
Λα γ Λβ δ η γδ = η αβ , α, β = 0, 1, 2, 3
and
Λα γ Λβ δ η αβ = η γδ , γ, δ = 0, 1, 2, 3
(physical motivation for and basic properties of the Lorentz group are dis-
cussed in some detail in the Introduction and first three sections of
Chapter 1 in [N3]).
ϕ̂ (y 0 , . . . , y 3 ) = ϕ (Λα 0 y α , . . . , Λα 3 y α ),
then ϕ̂ satisfies
( )(
∂ ∂
η αβ − i n Âα (y 0
, . . . , y 3
) − i nÂβ (y 0 ,
∂y α ∂y β
(2.3.8)
)
. . . , y 3 ) ϕ̂ (y 0 , . . . , y 3 ) + m2 ϕ̂ (y 0 , . . . , y 3 ) = 0.
( ) ( )
∂ ∂ ∂xβ
where A = Âα dy α so that Âα = A ∂y α = A ∂xβ ∂y α
=
( )
Λα β A ∂x∂ β = Λα β Aβ .
Remark: The essential, but rather camouflaged, issue here is that it is up to
us to specify what the wavefunction is to be in the new coordinates (i.e., how ϕ
transforms under L+↑ ) and that, in order to satisfy the requirements of special
relativity, we must do this in such a way that it satisfies “the same equation in
the new coordinate system.” In this case, the wavefunction transforms trivially
2.4. Spin One-Half Electrodynamics 71
Similarly,
( )( )
∂ ∂
− i nÂα (y , . . . , y )
0 3
− i nÂβ (y , . . . , y ) ϕ̂ (y 0 , . . . , y 3 )
0 3
∂y α ∂y β
( )( )
∂ ∂
γ
= Λα Λβ δ
− i nAγ (Λa y , . . .)
0 a
− i nAδ (Λa y , . . .)
0 a
∂xγ ∂xδ
× ϕ (Λa 0 y a , . . .).
and then refer those who are interested to the elegant, and quite accessible, ac-
count of these phenomena in Volume III of The Feynman Lectures on Physics
[Fey].
The classical Bohr picture of an atom (negatively charged electrons revolving
around a positively charged nucleus) suggests that an orbiting electron actu-
ally constitutes a tiny current loop. Such a current loop produces a magnetic
field which, at large distances, is the same as that of a magnetic dipole (lo-
cated at the center of the loop and perpendicular to the plane of the loop).
Such a dipole has a magnetic moment (a vector describing its orientation and
strength). Consequently, an electron in an atom has associated with it an “or-
bital magnetic moment.” Now, magnetic moments behave in interesting and
well-understood ways when subjected to external magnetic fields. In 1922 (just
before the advent of quantum mechanics), Stern and Gerlach carried out an
experiment designed to detect these effects for the orbital magnetic moment
of an electron. From the point of view of classical physics (the only point of
view available at the time), the results were quite shocking. Somewhat later,
the quantum mechanics of Schroedinger and Heisenberg, when applied to the
orbital magnetic moment of the electron, provided a qualitative, but not quan-
titative explanation of the outcome. Finally, it was suggested by Uhlenbeck
and Goudsmit that this discrepancy (and various others associated with the
anomalous Zeeman effect and the splitting of certain spectral lines) could
be accounted for if one assumed that the electron had associated with it an
additional magnetic moment, not arising from its orbital motion, but rather
from a sort of “intrinsic” angular momentum or “spinning” of the electron. The
suggestion was not that an electron actually spins on some axis in the same
way that the earth does on its, but rather that it possesses some intrinsic
property (called “spin”) that manifests itself in an external magnetic field by
mimicing the behavior of the magnetic moment of a spinning charged ball.
This intrinsic magnetic moment vector, however, must be of a rather peculiar
sort that one could only encounter in quantum mechanics. The Stern-Gerlach
experiment suggested that its component in any spatial direction could take
on only one of two possible values (±1 with the proper choice of units). This
has the following consequence. Let us select (arbitrarily) some direction in
space (say, the z-direction of some coordinate system). The intrinsic mag-
netic moment of an electron has, at each point, a z-component σz that can
take on one of the two values ±1. Which value it has will determine how the
electron responds to certain magnetic fields and so a complete description of
the electron’s wavefunction must contain this information. More precisely, the
wavefunction must be regarded as a function of not only x, y, z and t, but of
σz as well.
ψ = ψ(x, y, z, t, σz )
However, since σz can assume only the two values ±1, such a wavefunc-
tion is equivalent to a pair of functions ψ1 (x, y, z, t) = ψ (x, y, z, t, 1) and
ψ2 (x, y, z, t) = ψ (x, y, z, t, −1). It is convenient to put these two together into
2.4. Spin One-Half Electrodynamics 73
a column vector and adopt the point of view that an electron (or any spin
one-half particle) has a two-component wavefunction
( )
ψ1
ψ= .
ψ2
(here it is understood that both sides have been written in terms of one of
the two coordinate systems, x, y, z or x′ , y ′ and z ′ ). Assuming also (for the
moment) that the wavefunction is uniquely determined in each coordinate
system we find that a rotation by R2 ∈ SO(3) followed by a rotation by
R1 ∈ SO(3) must have the same effect as the rotation R1 R2 ∈ SO(3). Thus,
we must have
T (R1 R2 ) = T (R1 )T (R2 ).
Similarly, each T (R) must be invertible and satisfy
T (R−1 ) = (T (R))−1
What we find then is that the rule T which associates ) every R ∈ SO(3)
( with
ψ1
the corresponding transformation matrix T (R) for ψ2 is a homomorphism
into the group of invertible, 2 × 2, complex matrices. Identifying this latter
group with GL(C2 ) we find that T is a representation of SO(3) on C2 . The
reason this information is useful is that all of the representations of SO(3)
are known. These are usually described somewhat indirectly as follows: In
Appendix A of [N4] it is shown that SU (2) is the (double) covering group of
SO(3). More precisely, there exists a smooth, surjective group homomorphism
with kernel ± ( 10 01 ) and with the property that each point of SO(3) has an
open neighborhood V whose inverse image under Spin is a disjoint union of
(two) open sets in SU (2), each of which is mapped diffeomorphically onto V
by Spin. Now, consider a representation
h : SO(3) −→ GL(V)
SU(2)
Spin h=h
° Spin
SO(3) GL( )
h
( for V
k−r r
A basis )k consists of all polynomials z1 z2 , r = 0, 1, . . . , k. Each
g = αγ βδ in SU (2) gives rise to a linear transformation on Vk which
carries z1k−r z2r onto (z1′ )k−r (z2′ )r , where
( ) ( )( )
z1′ α γ z1
= .
z2′ β δ z2
and called the spin-j representation, where j = k2 . One can show (see
[vdW]) that each of these representations is irreducible (i.e., that there is
k
no proper subspace of Vk that is invariant under every D 2 (g), g ∈ SU (2))
and that every irreducible representation of SU (2) with complex representa-
tion space is equivalent to one of these (two representations D1 : G −→ GL(V1 )
and D2 : G −→ GL(V2 ) of a group G are equivalent if it is possible to choose
bases for V1 and V2 so that, for each g ∈ G, the matrices of D1 (g) and
D2 (g) are the same). Furthermore, any representation of SU (2) can be con-
structed from these irreducible representations by forming finite direct sums
(the direct sum of D1 : G −→ GL(V1 ) and D2 : G −→ GL(V2 ) is the repre-
sentation D1 ⊕ D2 : G −→ GL(V1 ⊕ V2 ) defined by (D1 ⊕ D2 )(g)(v1 , v2 ) =
(D1 (g)(v1 ), D2 (g)(v2 )). In effect, we now have all of the representations of
SU (2).
76 2. Physical Motivation
The polynomials have now served their purpose and it will be convenient to
note that Vk has complex dimension k + 1 and so can be identified with Ck+1
by identifying z1k−r z2r , r = 0, 1, . . . , k, with the standard basis for Ck+1 . The
k
linear transformations D 2 (g) can therefore be identified with (k + 1) × (k + 1)
complex matrices. For example, k = 0 gives the trivial representation of SU (2)
on C (D0 (g) = (1) for each g ∈ SU (2)), while k = 1 gives the identity
1
representation of SU (2) on C2 (D 2 (g) = g for every g ∈ SU (2)). Note that
1
D 2 is the only irreducible representation of SU (2) on C2 . The only other way
to get a representation of SU (2) on C2 is to form the direct sum of two copies
of D0 : ( )
1 0
(D ⊕ D )(g) = (1) ⊕ (1) =
0 0
.
0 1
1
This, of course, leaves everything in C2 alone. Finally notice that D 2 (−g) =
1
−D 2 (g) and (D0 ⊕ D0 )(−g) = (D0 ⊕ D0 )(g) so only this second example
descends to a representation of SO(3) (the trivial representation of SO(3) on
C2 ).
The situation we have just described would seem to present us with some-
thing of a dilemma. To ensure the rotational invariance of Pauli’s theory of
) SO(3) on C that would
2
the electron we were led to seek a representation
( of
ψ1
transform the two-component wavefunctions ψ2 when the coordinate sys-
tem is rotated. We find now that there is only one such (D0 ⊕ D0 ) and this
is the trivial
( ′ representation.
) Under
( this
) representation the transformed wave-
ψ
function ψ1′ would simply be ψ 1
ψ2 written in terms of the new coordinates.
2
However, this is clearly not consistent with the phenomenon (spin one-half)
which led us to two-component wavefunctions in the first place. Recall that
ψ1 (x, y, z) = ψ(x, y, z, 1) and ψ2 (x, y, z) = ψ(x, y, z, −1), where ±1 are the
possible z-components of the intrinsic magnetic moment of the electron. A
rotation which reverses the ( direction
) of the z-axis must interchange ψ1 and
ψ2 and so cannot leave ψ2 unchanged. Thus, D0 ⊕ D0 is not consistent
ψ1
with the structure we are attempting to model. Must we conclude then that
Pauli’s proposal is doomed to failure?
To extricate ourselves from this dilemma we must understand that there
is an essential feature of quantum mechanics that requires an adjustment
in the classical picture we painted earlier (pages 86–87). Our conclusion
that the transformation matrices T (R) satisfy T (R1 R2 ) = T (R1 )T (R2 ) and
(T (R))−1 = T (R−1 ) and therefore give rise to a representation of SO(3) fol-
lowed from the assumption that the wavefunction is uniquely determined in
each coordinate
( ) system. This, however, is not (quite) the case. For example,
both ± ψ ψ2
1
represent the same state of our electron since an overall sign
change has no effect on the probabilities described earlier (page 86) and all of
2.4. Spin One-Half Electrodynamics 77
SO(3)) that represents the physical process of rotating a frame through one
complete turn (360◦ ) about its x-axis (identify each element of SO(3) with
the configuration of the axes that would result from applying that rotation
to the initial configuration). The curve R2 (t) in SO(3) defined by the same
formula, but with 0 ≤ t ≤ 4π represents a rotation about the x-axis through
720◦ . Both R1 and R2 begin and end with the same configuration, but there
is a real difference, both physically and mathematically.
Spin : SU (2) −→ SO(3) is a covering space (Exercise A.13, [N4]) and cov-
ering spaces have the property that curves in the covered space lift uniquely to
curves in the covering space once an initial point is selected (Corollary 1.5.13,
[N4]).
ends at ( 10 01 ). Thus, a rotation of the frame through 360◦ changes the sign
of the wavefunction, but a rotation through 720◦ leaves the sign unchanged,
even though both rotations begin and end with the same configuration of the
axes. Mathematically, the difference between R1 and R2 is that they represent
two different homotopy classes in π1 (SO(3)) ∼ = Z2 . R2 is nullhomotopic since
it is Spin ◦g2 and g2 is a loop at ( 10 01 ) in SU (2) ∼
= S 3 , but R1 is not because
it lifts to a path g1 from ( 0 1 ) to − ( 0 1 ), in SU (2).
1 0 1 0
The next step in this program ( )would be to look at the differential equa-
tions proposed by Pauli for ψ 1
ψ2 and decide whether or not they assume
the same form when the coordinate system is rotated and the wavefunction is
1
transformed by D 2 (they do!). Since Pauli’s theory was eventually abandoned
(because it is not relativistically invariant) we shall not pursue this here, but
will instead turn to the profoundly successful alternative proposed by Dirac.
̸ ϕ = −i mϕ,
D (2.4.2)
80 2. Physical Motivation
i.e., if
γ α γ β = η αβ , α, β = 0, 1, 2, 3. (2.4.5)
Now, (2.4.5) clearly cannot be satisfied if the γ α are taken to be numbers
(none can be 0 since η αα = ±1, but γ α γ β = 0 if α ̸= β). Dirac’s idea was to
allow ϕ to have more than one complex component and interpret (2.4.5) as
matrix equations (one for each α, β = 0, 1, 2, 3). Specifically, if
ϕ1
.
ϕ= .
.
ϕn
be the Pauli spin matrices and taking σ0 to be the 2 × 2 identity matrix, one
easily verifies the usual commutation relations
σi2 = σ0 , i = 1, 2, 3
(2.4.7)
σi σj = −σj σi , i, j = 1, 2, 3, i ̸= j.
Now, we define
( ) 0 0 1 0
0 σ0 0 0 0 1
γ0 = =
σ0 0 1 0 0 0
0 1 0 0
( ) 0 0 0 −1
0 −σ1 0 0 −1 0
γ1 = =
σ1 0 0 1 0 0
1 0 0 0
( ) 0 0 0 i
0 −σ2 0 0 −i 0
γ2 = =
σ2 0 0 −i 0 0
i 0 0 0
( ) 0 0 −1 0
0 −σ3 0 0 0 1
γ3 = = .
σ3 0 1 0 0 0
0 −1 0 0
It is now a simple matter to verify that the conditions in (2.4.6) are satisfied
by these matrices, e.g.,
82 2. Physical Motivation
γ1γ2 + γ2γ1
( )( ) ( )( )
0 −σ1 0 −σ2 0 −σ2 0 −σ1
= +
σ1 0 σ2 0 σ2 0 σ1 0
( ) ( )
−σ1 σ2 0 −σ2 σ1 0
= +
0 −σ1 σ2 0 −σ2 σ1
( )
−(σ1 σ2 + σ2 σ1 ) 0
=
0 −(σ1 σ2 + σ2 σ1 )
1 0 0 0
( )
0 0 0 1 0 0
= [4pt] = 2η 12
0 0 1
0 0 0
0 0 0 1
and
( )( )
1 1 1 1 1 1 0 −σ1 0 −σ1
γ γ + γ γ = 2γ γ = 2
σ1 0 σ1 0
( ) ( )
−σ12 0 σ0 0
=2 = −2
0 −σ12 0 σ0
1 0 0 0
11 0 1 0 0
= 2η
0 0 1 0
0 0 0 1
etc.
Remark: There are many other possible choices for γ 0 , γ 1 , γ 2 and γ 3 ,
many of which are used in the physics literature. Indeed, for any nonsingular
matrix B, one can replace each γ α by Bγ α B −1 and obtain another set of “Dirac
matrices” satisfying (2.4.6). Conversely, one can show (see, e.g., pages 104–106
of [Gre]) that any set of 4 × 4 matrices satisfying (2.4.6) differs from our
choice by such a similarity transformation. Algebraically, this means that, up
to equivalence, there is only one representation of the Clifford algebra of R1,3
by 4 × 4 matrices. The choice we have made is called the Weyl, or chiral
representation.
With γ 0 , γ 1 , γ 2 and γ 3 the 4 ×4 matrices described above, the wavefunction
for our spin one-half particle has four components
2.4. Spin One-Half Electrodynamics 83
ϕ1
ϕ2
ϕ=
ϕ3
ϕ4
∂0 ∂0 ϕ = µi µj ∂i ∂j ϕ. (2.4.9)
84 2. Physical Motivation
(the reason for the peculiar numbering will become clear soon). One obtains an
operator D =(γ α ∂)α of the required type by taking, for example,
10
γ 0 = σ0 = 0 1 and γ i = µi = −σi , i = 1, 2, 3. Thus,
( ) ( ) ( ) ( )
1 0 0 −1 0 i −1 0
D= ∂ + ∂1 + ∂ + ∂
0 1 0 −1 0 −i 0 2 0 1 3
( )
∂0 − ∂3 −∂1 + i ∂2
= .
−∂1 − i ∂2 ∂0 + ∂3
and is known as the Weyl neutrino equation. Notice that this is just what
one would obtain from the Dirac equation with m = 0 and a wave- function
of the form
0
0
ϕ3 .
ϕ4
For future reference we note also that the m = 0 Dirac equation for a wave-
function of the form
ϕ1
ϕ
2
0
0
2.4. Spin One-Half Electrodynamics 85
reduces to ( )( ) ( )
∂0 + ∂3 ∂1 − i ∂2 ϕ1 0
= (2.4.13)
∂1 + i ∂2 ∂0 − ∂3 ϕ2 0
and that the coefficient matrix in (2.4.13) is the formal conjugate, transposed
inverse of the coefficient matrix in (2.4.12). The significance of these obser-
vations will emerge in our discussion of the transformation properties of the
wavefunctions and the corresponding invariance properties of the equations.
Now we return to the issue of the Lorentz invariance of the Dirac equation.
The problem, as it was for the Klein-Gordon equation (page 82), is to show
that the Dirac equation has the same form in any other coordinate system
y 0 , y 1 , y 2 , y 3 for R1,3 , related to the standard coordinates x0 , x1 , x2 , x3 by
y α = Λαβ xβ , α = 0, 1, 2, 3,
where Λ = (Λαβ ) ∈ L+↑ . However, since the Dirac wavefunction has four
complex components, this will require finding a representation T : L+↑ −→
GL(C4 ) of L+↑ on C4 which, if taken to be the transformation law for the
wavefunction, preserves the form of the Dirac equation (cf., the discussion
of the two-component Pauli theory on pages 86–93). As was the case for
SO(3) in the Pauli theory, it so happens that all of the representations L+↑
are known, that they are most conveniently described in terms of a two- fold
covering group of L+↑ and that, because of the nature of a quantum mechanical
wavefunction, it is actually the representations of the covering group that do
not descend to L+↑ that turn out to be of most interest. We begin with a brief
summary of the relevant results (see Chapter 3 for more details).
Identifying L+↑ with a subset of R16 one finds that it is a submanifold
diffeomorphic to SO(3) × R3 and so is a 6-dimensional Lie group containing
SO(3) as a closed subgroup. We denote by SL(2, C) the group of 2×2 complex
matrices with determinant one. Identifying SL(2, C) with a subset of C4 = R8
one finds that it is a submanifold diffeomorphic to S 3 × R3 and so it is also
a 6-dimensional (simply connected) Lie group. Note that SU (2) is a closed
subgroup of SL(2, C). Now, we have already described a two-fold covering
map
Spin : SU (2) −→ SO(3)
and we wish now to show that this is, in fact, the restriction to SU (2) of a
two-fold covering map of SL(2, C) onto L+↑ , also denoted
The construction of this map is carried out in detail in Section 1.7 of [N3]
so we will be brief. R1,3 can be identified with the linear space H of 2 × 2
86 2. Physical Motivation
= x0 σ0 + x1 σ1 + x2 σ2 + x3 σ3 = xα σα ,
where the squared norm is taken to be the determinant. Observe that each of
the coordinates xα can be expressed as
1
xα = trace(σα x)
2
so that ∑3
1
x= trace(σα x)σα .
α=0
2
Λg (x) = gxḡ ⊤ .
⊤
Note that Λg (x) is, indeed, in H because Λg (x) = (gxḡ ⊤ )⊤ = (ḡx̄g ⊤ )⊤ =
gxḡ ⊤ = Λg (x). Also note that, for g ∈ SU (2) ⊆ SL(2, C), ḡ ⊤ = g −1 . Now,
Λg is surely linear and satisfies det(Λg (x)) = det(gxḡ ⊤ ) = det(x) so it pre-
serves the Minkowski inner product on H = R1,3 . Thus, Λg is an orthogonal
transformation and, with a bit more work (page 77, [N3]), one can show that
it is proper and orthochronous. Now let
Λg (x) = y α σα .
Then
1 1
yα = trace(σα Λg (x)) = trace(σα gxḡ ⊤ )
2 2
1 β ⊤
= trace(σα g(x σβ )ḡ )
2
1
= trace(σα gσβ ḡ ⊤ )xβ
2
= Λαβ xβ
where
1
Λαβ = trace(σα gσβ ḡ ⊤ ), α, β = 0, 1, 2, 3.
2
Thus, (Λαβ )α,β=0,1,2,3 is in L+↑ and we define Spin: SL(2, C) −→ L+↑ by
for each g ∈ SL(2, C). One then shows that Spin is a two-fold covering group
for L+↑ and that its restriction to SU (2) agrees with the map of the same
2.4. Spin One-Half Electrodynamics 87
name discussed on page 88. Before proceeding we will record one additional
fact that we will need shortly. Notice that
1 1
Λαβ = trace(σα gσβ ḡ ⊤ ) = trace(σβ (ḡ ⊤ σα g))
2 2
so that
∑
3
ḡ ⊤ σα g = Λαβ σβ . (2.4.14)
β=0
Now we proceed just as we did for SO(3) in our discussion of the Pauli
theory. Consider a representation
h : L+↑ −→ GL(V)
of L+↑ . Composing with Spin then gives a representation of SL(2, C).
SL(2, C)
Spin h = h ° Spin
+ GL( )
h
defined by
( ) ( )( ) ( )
z1 α β z1 αz 1 + βz 2
−→ = .
z2 γ δ z2 γz 1 + δz 2
by
( )
( 12 ,0) (0, 12 ) g 0
D ⊕D (g) = ⊤ −1
,
0 (ḡ )
2.4. Spin One-Half Electrodynamics 89
where all of the entries are 2 × 2 matrices and we are again identifying a
linear transformation on C4 with its matrix relative to the standard basis.
Similarly, one can define the direct sum of any such pair. A somewhat less
obvious procedure for building a representation on C4 is the tensor product
of two representations on C2 . For instance, the representation
( )
can be described as follows: For each g = αγ βδ ∈ SL(2, C), we define
( )
( 12 , 12 ) α(ḡ ⊤ )−1 β(ḡ ⊤ )−1
D (g) =
γ(ḡ ⊤ )−1 δ(ḡ ⊤ )−1
αδ̄ −αγ̄ β δ̄ −βγ̄
−αβ̄ αᾱ −β β̄ β ᾱ
= .
γ δ̄ −γγ̄ δ δ̄ −δγ̄
−γ β̄ γ ᾱ −δ β̄ δ ᾱ
1 1 1 1 1 1
Note that D( 2 , 2 ) (−g) = D( 2 , 2 ) (g) so D( 2 , 2 ) descends to a representation of
L+↑ on C4 (also denoted D( 2 , 2 ) ). One can show that D( 2 , 2 ) is equivalent to
1 1 1 1
the natural (vector) representation of L+↑ on R1,3 . Similarly, one can define
1 1
such representations as D(1,0) = D( 2 ,0) ⊗ D( 2 ,0) . One can show that these
1 1
are irreducible, whereas such things as D( 2 ,0) ⊕ D(0, 2 ) , of course, are not.
From our point of view the important fact is that we have just described all
of the representations of SL(2, C) on C4 (up to equivalence). Proving the
relativistic invariance of the Dirac equation therefore amounts to searching
among these few representations of SL(2, C) on C4 for one which, if adopted
as the transformation law for the 4-component Dirac wavefunction, will lead
to a transformed wavefunction that satisfies the same (Dirac) equation in the
transformed coordinate system.
Begin with the Dirac equation in standard coordinates x = (x0 , x1 , x2 , x3 )
on R1,3 . ( )
γ β ∂β + i m ϕ(x) = 0. (2.4.15)
y α = Λαβ xβ , α = 0, 1, 2, 3,
∂ ∂y α ∂
∂β = = = Λαβ ∂ˆα .
∂xβ ∂xβ ∂y α
90 2. Physical Motivation
1 ( )
Λαβ = trace σα gσβ ḡ ⊤ , α, β = 0, 1, 2, 3.
2
Our objective is to find a representation
ρ : SL(2, C) −→ GL(C4 )
such that, if
ϕ̂(y) = ρ(g)(ϕ(Λ−1 y)),
then (2.4.15) implies
( )
γ α ∂ˆα + i m ϕ̂(y) = 0. (2.4.16)
γ β ∂β ϕ + imϕ = 0
( )( )
γ β Λαβ ∂ˆα (ρ(g))−1 ϕ̂(y) + im(ρ(g))−1 ϕ̂(y) = 0
( )
γ β (ρ(g))−1 Λαβ ∂ˆα ϕ̂(y) + (ρ(g))−1 imϕ̂(y) = 0.
This then is the condition that our representation ρ must satisfy in order to
preserve the form of the Dirac equation. We obtain a more convenient form
of this condition as follows:
we compute
( )
−1 0 0 g −1 σ0 (ḡ ⊤ )−1
(ρ(g)) γ ρ(g) = (2.4.19)
ḡ ⊤ σ0 g 0
and, for i = 1, 2, 3,
( )
−1 i 0 −g −1 σi (ḡ ⊤ )−1
(ρ(g)) γ ρ(g) = . (2.4.20)
ḡ ⊤ σi g 0
so
∑
3
0 Λα0 σ0 − Λαi σi
Λαβ γ β
= 3 i=1 . (2.4.21)
∑ Λα σ 0
β β
β=0
and, from a brief calculation that we will leave for the reader,
∑3
−1
0
− Λ0i σi ,
Λ σ
0 0 α =0
∑3
Λ β σβ =
α i=1
.
∑3
β=0 −Λα σ +
α
0 0 Λ i σi , α = 1, 2, 3
i=1
With this we have established the Lorentz invariance of the Dirac equation.
1 1
The emergence of the representation D( 2 ,0) ⊕ D(0, 2 ) as the appropriate
transformation law for a Dirac wavefunction has interesting and important
consequences that we will briefly explore. Let us write the Dirac wavefunction
ϕ as
ϕ1 ( )
ϕ2 ϕL
ϕ= = ,
ϕ3 ϕR
ϕ4
( ) ( ) 1
where ϕL = ϕ1
ϕ2 and ϕR = ϕ3
ϕ4 . Since ϕ transforms according to D( 2 ,0) ⊕
1 1 1
D(0, 2 ) , ϕL and ϕR transform according to D( 2 ,0) and D(0, 2 ) , respectively.
Furthermore, the Dirac equation becomes a pair of coupled equations for ϕL
and ϕR :
γ α ∂α ϕ = −imϕ
∑
3
0 σ0 ∂0 − σi ∂i ( ) ( )
ϕL ϕL
i=1 = −im
σ ∂ + ∑σ ∂
3
0 ϕR ϕR
0 0 i i
i=1
( )
∑3
σ0 ∂0 − σi ∂i ϕR = −imϕL
i=1
( ) . (2.4.22)
∑3
σ0 ∂0 + σi ∂i ϕL = −imϕR
i=1
which are, of course, just equations (2.4.12) and (2.4.13). To understand the
significance of ϕR and ϕL in general, the relationship between our current
model of spin one-half particles as 4-component objects and our earlier (2-
component) view of spin one-half (pages 85–86) and just what (2.4.23) and
(2.4.24) have to do with neutrinos (see page 100) we must discuss yet another
symmetry (invariance property) of the Dirac equation.
The Dirac equation is invariant under the proper, orthochronous Lorentz
group L+↑ because we were able to find a (“2-valued”) representation of L+↑
which, if taken to be the transformation law for the wavefunction, led to a
transformed wavefunction that satisfied the Dirac equation in the transformed
coordinate system. Now we wish to consider a coordinate transformation,
called spatial inversion, that does not correspond to an element of L+↑ . Its
matrix is
1 0 0 0
0 −1 0 0
π= 0
0 −1 0
0 0 0 −1
and its effect is simply to switch the orientation of the spatial coordinate sys-
tem. Note that, since π 2 is the 4 ×4 identity matrix, π generates a group of
coordinate transformations which we may denote Z2 , since that’s what it is
isomorphic to. In order to prove the invariance of the Dirac equation under
spatial inversions we will find a representation ρ : Z2 −→ GL(C4 ) which, if
taken to be the transformation law for ϕ, leads to a transformed wavefunction
that satisfies the Dirac equation in the transformed coordinate system. Note
that, since (γ 0 )2 is the 4 ×4 identity matrix, the assignments
id −→ id
π −→ γ 0
i.e., if
(ρ(Λ))−1 γ α ρ(Λ) = Λαβ γ β , α = 0, 1, 2, 3 (2.4.25)
(see page 107). Now, (2.4.25) is obviously satisfied if Λ = id so we need only
verify that it is also satisfied if Λ = π. In this case, (2.4.25) becomes
γ 0 γ α γ 0 = π αβ γ β , α = 0, 1, 2, 3,
i. e.,
γ 0γ 0γ 0 = γ 0
and
γ 0 γ i γ 0 = −γ i i = 1, 2, 3.
Since these are all easy to verify directly we have established the invariance
of the Dirac equation under spatial inversion. ( )
Notice, in particular, that if we write our Dirac wavefunction ϕ = ϕϕRL as
on page 109, then, under a spatial inversion, it transforms as follows:
( ) ( )( ) ( )
ϕL 0 σ 0 ϕL ϕR
ϕ= −→ γ 0 ϕ = = .
ϕR σ0 0 ϕR ϕL
(γ α ∂α + im)ϕ = 0
2.4. Spin One-Half Electrodynamics 95
obviously implies
(γ α ∂α + im)(eiθ ϕ) = 0.
The Dirac equation is therefore invariant under the global U (1)-action ϕ −→
eiθ ϕ. Just as was the case for the Schroedinger and Klein-Gordon equations,
elevating this global symmetry to a local gauge symmetry in which θ is a
function of x0 , x1 , x2 and x3 requires the presence of an electromagnetic gauge
potential. In the physics literature such potentials are included by “minimal
coupling,” i.e., by replacing the ordinary derivatives ∂α in the Dirac equation
by “covariant derivatives” ∂α + inAα (see pages 76–80).
The problem of molding all of the information we have assembled thus far
into a gauge theory model of the type described in Section 2.1 is complicated
by a number of issues. Recall that in the spin zero case (Section 2.2) we be-
gan with an electromagnetic field (i.e., a connection on a U (1)-bundle over
spacetime) and the various representations of U (1) on C and, from them,
constructed complex scalar fields, an action and the corresponding Euler-
Lagrange equations. The resulting Klein-Gordon equations happened to be
Lorentz invariant. Our point of departure in this section has been to insist
at the outset on the relativistic invariance of the free particle equations. This
led us to a wavefunction taking its values in C4 whose external symmetry
(Lorentz invariance) was expressed in the form of a transformation law corre-
sponding to a specific representation of the double cover SL(2, C) of L+↑ . This
suggests that, in a corresponding gauge theory model, the wavefunction is a
matter field on some SL(2, C)-bundle over spacetime. However, electrons are
coupled to electromagnetic fields and these are connections on U (1)-bundles,
not SL(2, C)-bundles, over spacetime. Furthermore, gauge invariance refers
specifically to the internal symmetry of a particle reflected in the behavior of
its wavefunction under changes in the local gauge potentials for the electro-
magnetic field so this notion also “lives” in a U (1)-bundle. To build a proper
gauge theory model for Dirac electrons coupled to electromagnetic fields will
require the “splicing together” of the external SL(2, C)-bundle and the in-
ternal U (1)-bundle into a single SL(2, C) × U (1)-bundle on which both the
electron and the electromagnetic field may be thought to live. This turns out
to be a relatively simple thing to do and we will outline the construction
shortly.
A more delicate, and much more interesting, obstacle is one that we could
evade altogether by simply continuing to restrict our attention to the space-
time R1,3 and its open submanifolds. From the perspective of the workaday
world of particle physics this would be an entirely reasonable choice since it
amounts to ignoring gravitational effects and these are generally negligible
in elementary particle interactions in the laboratory. From the perspective of
topology (and nonperturbative quantum field theory), however, such a choice
would “evade” the best part. We will conclude this section with a brief synop-
sis of the issues involved in describing spin one-half particles that live in more
96 2. Physical Motivation
general spacetimes where gravitational effects are not neglected. We will deal
with these issues in detail in the remaining chapters of the book.
A spacetime is a 4-dimensional (second countable, Hausdorff) manifold X
with a “Lorentz metric” g (this is a semi-Riemannian metric with the prop-
erty that each tangent space Tx (X) has a basis {e0 , e1 , e2 , e3 } for which
g(x)(eα , eβ ) = ηαβ ). Thus, each Tx (X) with its inner product g(x) can be
identified with R1,3 . The general Lorentz group L therefore acts on the or-
thonormal bases of each Tx (X). We are, however, only interested in bases
related by elements of L+↑ . Although one can isolate such a collection of bases
at each Tx (X) individually just by selecting an isomorphism onto R1,3 , an un-
ambiguous choice over the entire manifold X is possible only if X is assumed
orientable and “time orientable.” We will discuss this latter condition in more
detail in Chapter 3; essentially, one assumes the existence of a vector field V
on X that is timelike (g(x)(V (x), V (x))) > 0 for each x ∈ X) and so makes
a smooth selection over X of a timelike direction at each point that we may
(arbitrarily) decree “future-directed.” We will adopt both of these assump-
tions and thereby obtain, at each point, a family of oriented, time oriented,
orthonormal bases for the tangent space related by elements of L+↑ .
Now, Lorentz invariance means invariance under L+↑ . When X = R1,3 , bun-
dles over X are trivial so gauge fields, matter fields, etc., can all be identified
with objects defined on X. Furthermore, each tangent space can be canoni-
cally identified with X itself so L+↑ acts on X. In the general case, none of
this is true. In particular, choosing Lorentz frames and acting by L+↑ on such
frames cannot take place globally on all of X, but only point by point. To
describe all of this precisely we will build (in Chapter 3) the “oriented, time
oriented, orthonormal frame bundle” of X. This is a principal L+↑ -bundle
P
L+↑ ,→ L(X) −→
L
X
over X whose fibers consist of the oriented, time oriented, orthonormal bases
for the tangent spaces to X. The action of L+↑ on L(X) will simply carry one
such basis onto another (above the same point in X) and one can make sense
of “Lorentz invariance” for matter fields associated with this bundle by some
representation of L+↑ .
The frame bundle L+↑ ,→ L(X) −→ X exists for every oriented, time ori-
ented spacetime and so presents no real obstacle to our program. However,
there is an obstacle (or, rather, “obstruction”). The fibers of L(X) are all iso-
morphic to L+↑ , but the Dirac wavefunction did not arise from a representation
of L+↑ on C4 . Rather, it was determined by the representation D( 2 ,0) ⊕ D(0, 2 )
1 1
2.4. Spin One-Half Electrodynamics 97
SL (2, C)
Spin
The frame bundle provides a copy of L+↑ above every x ∈ X so what we need
is an SL(2, C)-bundle
P
SL(2, C) ,→ S(X) −→
S
X (2.4.26)
over X and a map of S(X) onto L(X) that is, in effect, the spinor map of
PS−1 (x) onto PL−1 (x) for each x ∈ X. More precisely, a spinor structure for
X consists of a principal SL(2, C)-bundle (2.4.26) over X and a map
λ : S(X) −→ L(X)
such that
PL (λ(p)) = PS (p)
and
λ(p · g) = λ(p) · Spin(g)
for all p ∈ S(X) and all g ∈ SL(2, C). The following diagram therefore
commutes.
s
S(X) × SL(2, C) S(X) X
λ × Spin λ idX
(X) × + (X) X
98 2. Physical Motivation
H(z, w) = z ⊤ γ 0 w,
where
( )
0 0 σ0
γ =
σ0 0
for all g ∈ SL(2, C). The required inner product on C4 is then given by
1
⟨z, w⟩ = (H(z, w) + H(w, z)).
2
2.4. Spin One-Half Electrodynamics 99
ϕ : S(X) −→ C4
Let
P1 ◦ P2 = {(p1 , p2 ) ∈ P1 × P2 : P1 (p1 ) = P2 (p2 )}.
by
P12 (p1 , p2 ) = P1 (p1 ) = P2 (p2 ).
Then P12 is a smooth map of P1 ◦ P2 onto X. Define a smooth right action
of G1 × G2 on P1 ◦ P2 by
Then
P
G1 × G2 ,→ P1 ◦ P2 −→
12
X
is a smooth principal G1 × G2 -bundle over X.
Next we define maps π1 : P1 ◦ P2 −→ P1 and π2 : P1 ◦ P2 −→ P2 by
πi (p1 , p2 ) = pi , i = 1, 2. Letting e1 and e2 denote the identities in G1 and G2
we identify {e1 } × G2 with G2 and G1 × {e2 } with G1 . We then have principal
bundles
π1
G2 ,→ P1 ◦ P2 −→ P1
and
π
G1 ,→ P1 ◦ P2 −→
2
P2
for which the following diagram commutes:
P1 ° P2
π1 π2
P1 12 P2
1 2
X
ρ1 × ρ2 : G1 × G2 −→ GL(V)
by
P
Now, a matter field on G1 × G2 ,→ P1 ◦ P2 −→
12
X associated with ρ1 × ρ2 :
G1 × G2 −→ GL(V) is a map ϕ : P1 ◦ P2 −→ V satisfying
i.e., ( ( )) (( ))
ϕ ((p1 · g1 , p2 · g2 )) = ρ1 g1−1 ρ2 (g2−1 (ϕ (p1 , p2 ))).
Now we apply this construction to the following special case. Begin with an
oriented, time oriented spacetime X and a spinor bundle
P
SL(2, C) ,→ S(X) −→
S
X
P
on X (this will be G1 ,→ P1 −→ 1
X). Let ω 1 be the spinor connection on
S(X) referred to in the Remark on page 117. Take V = C4 and let ρ1 be the
representation D( 2 ,0) ⊕ D(0, 2 ) of SL(2, C) on C4 . Thus,
1 1
z1 ( ) z1
. g1 0 .
(ρ1 (g1 )) .
. = 0 (ḡ ⊤ )−1 .
.
1
z4 z4
P
for each g1 in SL(2, C). Next let U (1) ,→ P −→ 2
X be some princi-
pal U (1)-bundle over X and let ω 2 be a connection on it (representing
102 2. Physical Motivation
some electromagnetic field to which the Dirac electron will respond). Take
ρ2 : U (1) −→ GL(C4 ) to be the representation given by
z1 z1 g2 z1
. . .
(ρ2 (g2 )) . . .
. = g2 . = .
z4 z4 g2 z4
for each g2 ∈ U (1). Note that ρ1 (g1 ) ◦ ρ2 (g2 ) = ρ2 (g2 ) ◦ ρ1 (g1 ) as required in
(2.4.27). Thus, we have a representation
ρ1 × ρ2 : SL(2, C) × U (1) −→ GL(C4 )
given by
zl z1
. ..
(ρ1 × ρ2 )(g1 , g2 ) .
. = ρ1 (g1 ) ◦ ρ2 (g2 ) .
z4 z4
( ) g2 z1
g1 0 .
= ( ⊤ )−1 .
.
0 ḡ1
g2 z4
( ) z1
g1 0 .
= g2 ( ⊤ )−1 .. .
0 ḡ1
z4
i.e.,
( )
(g1 g2 )−1 0
ϕ ( p1 · g1 , p2 · g2 ) = ⊤ ϕ ( p1 , p2 ),
0 (g1 g2 )
particle with an electromagnetic field (the details are available in Section 7.2
of [Bl]).
As one final illustration of this technique we will sketch an analogous con-
struction for the interaction of a nucleon with a classical Yang-Mills field (for
the details, see Section 7.3 of [Bl]). Here we face the same problem as in the
case of a Dirac electron coupled to an electromagnetic field. A Yang-Mills
field is given by a connection on a principal SU (2)-bundle over spacetime
(Section 6.3 of [N4]), whereas a nucleon (proton/neutron) is a massive, spin
one-half particle and therefore lives in a spinor bundle. There is an additional
complication, however. A nucleon is a proton/neutron doublet, i.e., its wave-
function has a proton component and a neutron component and so must take
its values in V = C4 ⊕ C4 = C8 .
We begin then with an oriented, time oriented spacetime X and a spinor
bundle
PS
SL(2, C) ,→ S(X) −→ X.
ω 1 is again the spinor connection referred to in the Remark on page 117. Now
take V = C4 ⊕ C4 , which we identify with the set of ( vv12 ) with v1 , v2 ∈ C4 .
Letting ρ = D( 2 ,0) ⊕ D(0, 2 ) we define ρ1 : SL(2, C) −→ C4 ⊕ C4 by
1 1
( ) ( )
v1 (ρ(g1 ))(v1 )
( ρ1 ( g1 ))(v) = ( ρ1 ( g1 )) = .
v2 (ρ(g1 ))(v2 )
P
Now let SU (2) ,→ P −→ 2
X be some principal SU (2)-bundle over X and
ω 2 some connection on it (representing the Yang-Mills potential to which
the nucleon is coupled). Define ρ2 : SU (2) −→ GL(C4 ⊕ C4 ) as follows: For
each
( )
α β
g2 =
γ δ
in SU (2),
( ) ( )( )
v1 α β v1
( ρ2 ( g2 ) )(v) = ( ρ2 ( g2 ) ) =
v2 γ δ v2
( )
αv1 + βv2
= .
γvl + δv2
Letting
z1
..
( ) .
v1
v= = z4
v2 w1
.
.
.
w4
we have ( )
( ρ1 × ρ2 ( g1 , g2 ) )(v) = ( ρ( g1 ) ) (αv 1 + βv 2 )
( ρ( g1 ) ) (γv1 + δv2 )
( )
αz1 + βw1
g1 ( 0) ..
−1 .
0 ḡ1⊤
αz4 + βw4
= .
( )
γz1 + δw1
g1 ( 0) ..
−1 .
0 ḡ1⊤
γz4 + δw4
ϕ ( ( p1 , p2 ) · ( g1 , g2 ) ) = ( g1−1 , g2−1 ) · ϕ( p1 , p2 ).
( )
Writing ϕ = ϕϕ12 we have
( ρ (g1−1 ))(αϕ1 (p1 · g1 , p2 · g2 )
( )
ϕ1 ( p1 · g1 , p2 · g2 ) +βϕ2 ( p1 · g1 , p2 · g2 ) )
= .
ϕ2 ( p1 · g1 , p2 · g2 ) (ρ (g −1 )) (γϕ ( p · g , p · g )
1 1 1 1 2 2
+δϕ2 ( p1 · g1 , p2 · g2 ) )
for all g ∈ SU (2) and A ∈ su(2). Note that ⟨adg (A), adg (B)⟩ =
⟨gAg −1 , gBg −1 ⟩ = −2 trace((gAg −1 )(gBg −1 )) = −2 trace(g(AB)g −1 ) = −2
trace(AB) = ⟨A, B⟩, as required. Since every bundle over Rn is trivial, we
will trivialize at the outset and take
P
SU (2) ,→ Rn × SU (2) −→ Rn
for all p = (x, h) ∈ Rn × SU (2) and all g ∈ SU (2). There is a natural global
cross-section s : Rn −→ Rn × SU (2) given by
s(x) = (x, e)
where, for convenience, we write e for the identity element in SU (2). Any
other global cross-section then has the form
sg : Rn −→ Rn × SU (2)
sg (x) = s(x) · g(x)
= (x, e) · g(x)
= (x, g(x))
for some smooth map g : Rn −→ SU (2) (Exercise 4.3.5 of [N4]). The cross-
section sg gives rise to an automorphism of the bundle (i.e., a global gauge
transformation) in the usual way (see page 343 of [N4]):
s(x) · h −→ sg (x) · h
(x, e) · h −→ (x, g(x)) · h
(x, h) −→ (x, g(x)h).
106 2. Physical Motivation
Ag = (sg )∗ ω
related to A by
Ag = g −1 Ag + g −1 dg,
where dg is the entrywise exterior derivative of g : Rn −→ SU (2) and the
products are matrix products (we will do an explicit calculation of this sort
for the “t’ Hooft-Polyakov monopole” somewhat later). The curvature Ω of ω
is likewise uniquely determined by the field strength
1
F = s∗ Ω = dA + A ∧ A = Fαβ dxα ∧ dxβ ,
2
where
Fαβ = ∂α Aβ − ∂β Aα + [Aα , Aβ ], α, β = 1, . . . , n.
A gauge transformation g : Rn −→ SU (2) gives a new field strength
F g = (sg )∗ Ω
related to F by
F g = g −1 F g.
In this context a matter field is a smooth su(2)-valued map Φ on
Rn × SU (2) that satisfies
Φ( p · g) = g −1 · Φ(p)
Φ( (x, h) · g) = adg−1 (Φ(x, h))
Φ (x, hg ) = g −1 Φ(x, h)g
for all (x, h) ∈ Rn × SU (2) and all g ∈ SU (2). When, as in this case, V is the
Lie algebra G of the structure group G and ρ is the adjoint representation of
G on G, a matter field is referred to as a Higgs field. The triviality of the
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 107
bundle in our present circumstances allows us to identify the Higgs field with
its pullback by the global cross-section s:
ϕ = s∗ Φ = Φ ◦ s
ϕ(x) = Φ(x, e).
Under a gauge transformation g : Rn −→ SU (2),
ϕg = (sg )∗ Φ = g −1 ϕg
because
((sg )∗ Φ)(x) = Φ(sg (x))
= Φ(x, g)
= Φ(x, eg)
= g −1 Φ(x, e)g
= g −1 ϕ(x)g.
The next item on the agenda (#7 of Section 2.1) is the potential function
U : su(2) −→ R. This plays a rather peculiar role in the story we wish
to tell. Initially we adopt what is called the Georgi-Glashow potential
U : su(2) −→ R given by
λ
U (A) = (||A∥2 − 1)2 ,
8
where λ ≥ 0 is a constant and ∥ A ∥2 = ⟨ A, A ⟩ = −2 trace(A2 ), noting that
U (g · A) = U (gAg −1 ) = U (A) as required. Shortly, however, we will take λ to
be zero and retain only a vestige of the potential in the form of an asymptotic
boundary condition that it imposes on the Higgs field ϕ (see pages 132–133).
In order to describe the appropriate action (#8 of Section 2.1) for our exam-
ple we must anticipate a few results on differential forms that will be proved
later (Chapter 4). We will content ourselves with just a brief description of
those particular items required for the example. We have already introduced
real- and vector-valued 0-forms (pages 10 and 12), 1-forms (pages 9 and 12),
and 2-forms (pages 11 and 12). k-forms, for integers k ≥ 3, are defined analo-
gously and all of the familiar algebraic and analytic operations on 0-, 1-, and
2-forms extend to this more general context. For example, a real-valued 3-form
on a manifold X is a map α that assigns to each p ∈ X a real-valued trilin-
ear function αp on Tp (X) × Tp (X) × Tp (X) that is skew-symmetric (changes
sign when-ever two of its arguments are interchanged) and smooth in the
sense that, for any V 1 , V 2 , V 3 ∈ X (X), the function α(V 1 , V 2 , V 3 ) on
X defined by (α(V 1 , V 2 , V 3 ))(p) = αp (V 1 (p), V 2 (p), V 3 (p)) is in C ∞ (X).
These arise, for example, as exterior derivatives of 2-forms and wedge prod-
ucts of 1-forms and 2-forms, or of three 1-forms (all of which will be defined
carefully in Chapter 4).
In general, the set of k-forms on an n-dimensional manifold X is denoted
Λk (X) and admits a natural C ∞ (X)-module structure (and so in particular,
is a real vector space). If (U, φ) is a chart for X with coordinate functions
108 2. Physical Motivation
called the Hodge star operator. Moreover, dim Λn (X) = ( nn ) = 1 and, when
X is oriented and has a metric, there is a distinguished generator for Λn (X)
called the metric volume form and denoted vol (in standard coordinates on
Rn this is just dx1 ∧ · · · ∧ dxn ). In particular, for any α, β ∈ Λk (X), α ∧ ∗β is
in Λn (X) and so is a multiple, by some element of C ∞ (X), of vol. We denote
this element of C ∞ (X) by ⟨α, β ⟩:
α ∧∗ β = ⟨α, β ⟩vol.
and
( ) ( ) ( )
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 =
1 0 i 0 0 −1
are the Pauli spin matrices. It is easy to see that {T1 , T2 , T3 } is orthonormal
with respect to the inner product ⟨ A, B ⟩ = −2 trace(AB) on su(2). Then
any su(2)-valued k-form φ on Rn (e.g, A or F ) can be regarded as a matrix
of complex k-forms
( )
3 2 1
1 1 φ i φ + φ i
φ = φa Ta = − φa (i σa ) = − (2.5.1)
2 2 −φ2 + φ1 i −φ3 i
1
φ= φi ···i dxi1 ∧ · · · ∧ dxik , (2.5.2)
k! 1 k
where
the complex wedge product described above. We will illustrate the procedure
with an example that will also allow us to write out our action functional. We
consider a k -form φ with values in su(2) and written in the form (2.5.1) and
will compute φ ∧ ∗ φ, where ∗ φ is the Hodge dual of φ, computed componen-
twise (i.e., entrywise). Thus,
( )( )
∗ 3 ∗ 2
∗ 1 φ3 i φ2 + φ1 i φ i φ + ∗ φ1 i
φ∧ φ=
4 −φ2 + φ1 i −φ3 i −∗ φ 2 + ∗ φ 1 i −∗ φ3 i
−φ3 ∧ ∗ φ3 (φ3 i ) ∧ (∗ φ2 + ∗ φ1 i )
+(φ2 + φ1 i ) −(φ2 + φ1 i )
1 ∧(−∗ φ2 + ∗ φ1 i ) ∧(∗ φ3 i )
= .
4
(−φ2 + φ1 i ) (−φ2 + φ1 i )
∧(∗ φ3 i ) − (φ3 i ) ∧(∗ φ2 + ∗ φ1 i )
∧(−∗ φ2 + ∗ φ1 i ) −φ2 ∧ ∗ φ3
and similarly for the rest. In particular, the (2, 2)-entry is the same so
We define
∥φ∥2 = ∥φ1 ∥2 + ∥φ2 ∥2 + ∥φ3 ∥2
so that
−2trace(φ ∧ ∗ φ) = ∥φ∥2 vol. (2.5.4)
With the machinery we have assembled thus far we can write out the Yang-
Mills-Higgs action functional A(A, ϕ) for our example:
∫ (
A(A, ϕ) = − trace(F ∧ ∗ F ) − trace(dA ϕ ∧ dA ϕ)
Rn
λ∗ )
+ (∥ϕ∥2 − 1)2
8
∫ ( (2.5.5)
1
= ∥F ∥2 + ∥dA ϕ∥2
2 Rn
λ )
+ (∥ϕ∥2 − 1)2 dx1 ∧ · · · ∧ dxn ,
4
where dA ϕ = dϕ + [A, ϕ] is the covariant exterior derivative of the Higgs field
ϕ. We observe first that A(A, ϕ) is gauge invariant, i.e., that the effect of a
gauge transformation g : Rn −→ SU (2) (A −→ Ag , F −→ F g and ϕ −→ ϕg )
is to leave the integral unchanged. We have already seen that F g = g −1 F g
and ϕg = g −1 ϕg and we show now that
( )
dA ϕg = g −1 dA ϕ g
g
(2.5.6)
as well (we will need to use a few simple algebraic properties of d, e.g., the
product rule for matrix products, that will be proved in Chapter 4). Indeed,
dA ϕ = dϕ + [A, ϕ] implies
( )
g −1 d A ϕ g = g −1 dϕg + g −1 [A, ϕ]g
and
( ) [ ]
d A ϕg = dϕg + [Ag , ϕg ] = d g −1 ϕg + g −1 Ag + g −1 dg, g −1 ϕg
g
( ) [ ] [ ]
= d g −1 ϕg + g −1 Ag, g −1 ϕg + g −1 dg, g −1 ϕg
( ) [ ]
= g −1 [A, ϕ]g + d g −1 ϕg + g −1 dg, g −1 ϕg
because
g −1 g = id =⇒ g −1 dg + dg −1 g = 0
=⇒ g −1 dg = −dg −1 g
=⇒ g −1 dg g −1 ϕg = −dg −1 ϕg.
Gauge invariance of the action will therefore follow if we can show that
∥g −1 φg∥2 = ∥φ∥2 for any su(2)-valued form φ. But if we write φ = φa Ta ,
then g −1 φg = φa (g −1 Ta g) and, since ⟨g −1 Ag, g −1 Bg⟩ = ⟨A, B⟩, {g −1 T1 g,
g −1 T2 g, g −1 T3 g} is also an orthonormal basis for su(2) so this is clear (see the
Remark on page 130).
We are interested in finite action, stationary configurations (A, ϕ), i.e.,
solutions to the Euler-Lagrange equations for the action (2.5.5) for which
A(A, ϕ) < ∞. As it happens, no such solutions exist when n > 4 (see [JT]).
When n = 2 such solutions do exist and they are called vortices. These are
studied exhaustively in [JT], but we will have no more to say about them.
When n = 4 any such solution is gauge equivalent to a pure Yang-Mills field
(λ = 0, ϕ = 0) of the type discussed in [N4] (we will briefly review this
material shortly). Our primary concern is with the case n = 3 where finite
action, stationary configurations are, for reasons we hope to make clear, called
monopoles. In fact, we intend to discuss only a special case in which just
a vestige of the Georgi-Glashow potential survives. This special case arises
in the following way: The requirement that A(A, ϕ) < ∞ implies that, as
|x| −→ ∞ in R3 ,
F −→ 0 (2.5.7)
dA ϕ −→ 0 (2.5.8)
and, at least if λ ̸= 0,
ϕ −→ 1. (2.5.9)
The configuration (A, ϕ) must also satisfy the following Bianchi identities:
{
dA F = 0
. (2.5.13)
dA dA ϕ = [F , ϕ]
Thus we are looking for solutions to (2.5.12) that live in C. We shall find
some interesting ones, but not by studying (2.5.12) directly. As it happens,
there is a simpler set of first order equations whose solutions necessarily also
satisfy (2.5.12) and, in fact, give the absolute minima of the action (2.5.10).
Remark: This is entirely analogous to the situation encountered in [N4],
where the (anti-) self-dual equations on R4 gave the absolute minima of the
Yang-Mills action. Shortly we will review that situation and find that there is
a closer connection than simple analogy.
The best way to see where these equations come from is as follows: We denote
by ⟨ , ⟩ the inner product we have defined on su(2)-valued forms on R3
(Remark, page 130). Notice that, on R3 , F and ∗ dA ϕ are both 2- forms and
(since the metric on R3 is Riemannian), ∥dA ϕ∥2 = ∥∗ dA ϕ∥2 . Now notice that
2 2
∗ A
∥F ∥2 + dA ϕ = ∥F ∥2 + d ϕ
⟨ ⟩
= ⟨F , F ⟩ + ∗ dA ϕ, ∗ dA ϕ
⟨ ⟩ ⟨ ⟩
= F − ∗ dA ϕ, F − ∗ d A ϕ + 2 F , ∗ dA ϕ
2 ⟨ ⟩
= F − ∗ dA ϕ + 2 F , ∗ dA ϕ
and, similarly,
2 2 ⟨ ⟩
∥F ∥2 + d A ϕ = F + ∗d Aϕ − 2 F , ∗d Aϕ .
F = ±∗ dA ϕ. (2.5.14)
dA ∗ F = 0, (2.5.17)
dA F = 0. (2.5.18)
Now notice that if we have a potential A for which the field strength F is
self-dual (SD)
∗
F = F, (2.5.19)
or anti-self-dual (ASD)
∗
F = −F , (2.5.20)
then the Bianchi identity (2.5.18) implies that F necessarily satisfies the Yang-
Mills equations (2.5.17). Such potentials A are called SU (2) instantons on
R4 and, in [N4], a number of examples (called the BPST instantons) were
described. Written in quaternionic notation (i.e., identifying R4 with H and
su(2) with the Lie algebra Im H of pure imaginary quaternions) these can be
written ( )
q̄ − n̄
Aλ,n (q) = Im dq , (2.5.21)
λ2 + |q − n|2
where λ > 0 and n ∈ H are parameters called the scale and center of the
instanton, respectively. The corresponding field strengths are
λ2
F λ,n (q) = dq̄ ∧ dq. (2.5.22)
(λ2 + |q − n|2 )2
116 2. Physical Motivation
over S 4 up to equivalence.
Choose any connection ω on the bundle and let Ω denote its curvature
(we will prove in Chapter 3 that connections exist on any smooth principal
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 117
and called the second Chern number of the bundle (physicists call
−c2 (P )[S 4 ] the topological charge or instanton number of the bundle).
The 8π1 2 ensures that c2 (P )[S 4 ] is an integer (not obvious, but true) and it
is this integer that labels the equivalence classes of SU (2)-bundles over S 4 .
More precisely, two principal SU (2)-bundles over S 4 are equivalent if and only
if their second Chern numbers are equal.
118 2. Physical Motivation
Remark: We have already seen that all of the BPST potentials Aλ,n are
ASD and have Yang-Mills action 8 π 2 . They must, of course, have the same
Yang-Mills action since they all extend to the same bundle, i.e., the Hopf
bundle (which we now see has Chern number 1).
In order to establish contact with the Bogomolny monopole equations we
consider an arbitrary potential  = Âα dxα on R4 with field strength F̂ =
2 F̂αβ dx ∧ dx and use (6.4.4) of [N4] to write the components of the Hodge
1 α β
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 119
dual ∗ F̂ as
1 ∑
4
∗
F̂ αβ = ϵαβγδ F̂αβ , α, β = 1, 2, 3, 4,
2
γ,δ=1
1 ∑
4
F̂αβ = ∓ ϵαβγδ F̂αβ , α, β = 1, 2, 3, 4. (2.5.30)
2
γ,δ=1
Using i, j and k for indices taking the values 1, 2 and 3 one finds (by just
writing them out) that all of these equations are contained in the following:
∑
3
F̂ij = ± ϵijk F̂k4 , i, j = 1, 2, 3, (2.5.31)
k=1
where ϵijk is totally anti-symmetric in ijk and ϵ123 = 1. For example, taking
α = 3 and β = 4 in (2.5.30) gives
1 ∑ 1[ ]
4
F̂34 = ∓ ϵ34γδ F̂γδ = ∓ ϵ3412 F̂12 + ϵ3421 F̂21
2 2
γ,δ=1
1 [ ]
=∓ 2ϵ3412 F̂12 = ∓ ϵ3412 F̂12 = ± ϵ4123 F̂12
2
= ± ϵ123 F̂12
which is equivalent to
F̂12 = ± ϵ123 F̂34
of x4 , (2.5.31) becomes
∑
3
F̂ij = ± ϵijk F̂k4
k=1
∑3 ( [ ])
=± ϵijk ∂k Â4 − ∂4 Âk + Âk , Â4
k=1
∑3 ( [ ])
F̂ij = ± ϵijk ∂k Â4 + Âk , Â4 , i, j = 1, 2, 3 (2.5.32)
k=1
(you may wish to glance back at (2.5.15) if you’re wondering where all of this
is going).
Now we “reduce to R3 ” as follows: Fix some value x40 of x4 and consider the
submanifold R3 ×{x40 } of R4 (which we henceforth identify with R3 ). Restrict
our trivial SU (2)-bundle over R4 to this R3 and obtain a trivial SU (2)-bundle
over R3 . Let A1 , A2 and A3 be the restrictions to R3 of Â1 , Â2 and Â3 (all
of which are assumed independent of x4 ). Then
∑
3
Fij = ± ϵijk (∂k ϕ + [Ak , ϕ]), i, j = 1, 2, 3. (2.5.33)
k=1
have (for reasons that no doubt remain obscure) called a monopole. To under-
stand the terminology and to see Dirac monopoles in an entirely new light we
will describe now the t’Hooft-Polyakov-Prasad-Sommerfleld monopole,
which is an exact solution to the equations (2.5.33). For this we will need some
notation. First, in R3 we will write
x = (x1 , x2 , x3 )
√
r = |x| = (x1 )2 + (x2 )2 + (x3 )2
ni = xi /r, i = 1, 2, 3 (r ̸= 0)
d⃗x = (dx1 , dx2 , dx3 )
and, in su(2),
1
Ta = − i σa , a = 1, 2, 3
2
T⃗ = (T1 , T2 , T3 ).
In addition we set
⃗n · T⃗ = n1 T1 + n2 T2 + n3 T3 = na Ta
⃗n × T⃗ = (n2 T3 − n3 T2 , n3 T1 − n1 T3 , n1 T2 − n2 T1 )
and
( )
⃗n × T⃗ · d⃗x = (n2 T3 − n3 T2 ) dx1 + (n3 T1 − n1 T3 ) dx2
+ (n1 T2 − n2 T1 ) dx3 .
Our objective is to find (A(x), ϕ(x)) which satisfies (2.5.33) and has the re-
quired asymptotic behavior (∥ϕ∥ −→ 1 as r −→ ∞).
Although the monopole equations (2.5.33) are substantially less complicated
that the full Yang-Mills-Higgs equations, they are still far beyond the means
of elementary techniques. To reduce the level of difficulty a bit more requires
a guess (physicists prefer the term Ansatz) as to the form one might expect
for a solution. Here’s the one that worked for t’Hooft and Polyakov (some
rationale for the Ansatz is discussed in Section 4.2 of [GO]): We will seek
functions f (r) and h(r) satisfying
f (r)
−→ 1 as r −→ ∞ and f (0) = 0, (2.5.34)
r
and
h (r) −→ 0 as r −→ ∞ and h(0) = 1, (2.5.35)
such that
f (r) ( ⃗ ) f (r) a
ϕ (x) = ⃗n · T = 2 x Ta (2.5.36)
r r
122 2. Physical Motivation
and
1 − h(r) ( )
A (x) = ⃗n × T⃗ · d⃗x
r
(2.5.37)
1 − h(r) ∑
3
= ϵaij xj dxi Ta
r2 a=1
F 2 − H 2 = 1.
To get solutions f and h satisfying f (0) = 0 and h(0) = 1 we take for F and
H satisfying F 2 − H 2 = 1 the following:
F (r) = − coth r and H(r) = csch r.
Then
f (r) = r coth r − 1 and h(r) = r csch r
and one can verify directly that these are smooth (even at r = 0, where
they take the required boundary values) and satisfy (2.5.38). Thus, our field
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 123
Remark: The same calculation for (2.5.33) with the plus sign gives the
same result except that the sign of ϕ is changed.
The exact solution (A, ϕ) given by (2.5.39) and (2.5.40) is remarkable for
a number of reasons (quite aside from the fact that it is an exact solution
which is more than one generally has a right to expect). The form of the thing
itself is extraordinary for its “mixing” of the spatial and internal directions.
For example, (2.5.39) describes a Higgs field which, in the xa -direction in
R3 , a = 1, 2, 3, has only an internal T a -component. It is, in some strange way,
“radial” (Polyakov called it a “hedgehog” solution). Another feature, and one
that will be particularly significant quite soon, is that, despite appearances,
the component functions of ϕ and A are smooth (in fact, real analytic), even
at the origin. For example,
( ) ( )
1 1 r cosh r 1 3
1 r + 2! r + 4!1 5
r + ···
coth r − = −1 = −1
r r sinh r 1 3
r r + 3! r + 5!1 5
r + ···
( )
1 1 + 2! 1 2
r + ···
1 4
r + 4!
= − 1
r 1 + 3! 1 2
r + ···
1 4
r + 5!
( ( ) )
1 1 1
= 1+ − r + ··· − 1
2
r 2! 3!
(by long division) and this is indeed analytic at r = 0. Computing the deriva-
tive of coth r− 1r one finds that it is positive. Moreover, limr−→∞ (coth r− 1r ) =
1 so one obtains the following picture of ∥ϕ∥:
|| φ(r) || = coth r – 1r
0 r
ϕ −→ ϕg = g −1 ϕg
A −→ Ag = g −1 Ag + g −1 dg.
φ
−αβ̄ − αβ = −α(β̄ + β) = − cos (2Re(β))
2
φ φ x1
= 2 cos sin cos θ = sin φ cos θ =
2 2 r
φ φ
α2 − β 2 = cos2 − e−2i θ sin2
2 2
φ φ
− (cos 2θ − i sin 2θ) sin2
= cos2
2 2
( φ φ ) φ
= cos2 − cos 2θ sin2 + i sin 2θ sin2
2 2 2
( ) ( )
i αβ̄ − αβ = i α β̄ − β = i α(−2i Im(β))
φ φ
= 2α Im (β) = 2 cos (sin θ sin )
2 2
x2
= sin φ sin θ =
r
φ φ φ
−α2 − β 2 = (− cos2 − cos 2θ sin2 ) + i sin 2θ sin2
2 2 2
φ φ
α2 − β β̄ = cos2 − sin2
2 2
x3
= cos φ =
r
φ ( −i θ φ)
2αβ = 2 cos −e sin = −e−i θ sin φ
2 2
= − cos θ sin φ + i sin θ sin φ.
We need just the (1,1) and (1,2) entries of ϕg so we compute as follows: The
(1,1) entry of xa (g −1 σa g) is
φ φ φ
x1 (cos2 − cos 2θ sin2 ) + x2 (− sin 2θ sin2 ) + x3 (− cos θ sin φ)
2 2 2
2 φ φ
= (r sin φ cos θ) cos − (r sin φ cos θ) cos 2θ sin2
2 2
2 φ
− (r sin φ sin θ) sin 2θ sin − r cos φ cos θ sin φ
( )2
1 1
= r sin φ cos θ + cos φ
2 2
( )
1 1
− r sin φ cos θ(1 − 2 sin θ)2
− cos φ
2 2
( )
1 1
− 2r sin φ sin θ cos θ
2
− cos φ − r cos φ cos θ sin φ
2 2
1 1
= r sin φ cos θ + r sin ϕ cos φ cos θ
2 2 ( )
1 1
− (r sin φ cos θ − 2r sin φ cos θ sin θ)
2
− cos φ
2 2
( )
1 1
− 2r sin φ cos θ sin2 θ − cos φ − r cos φ cos θ sin φ
2 2
1 1
= r sin φ cos θ + r sin φ cos φ cos θ
2 2( )
1 1
− (r sin φ cos θ) − cos φ − r sin φ cos φ cos θ
2 2
1 1
= r sin φ cos θ − r sin φ cos φ cos θ
2 2
1 1
− r sin φ cos θ + r sin φ cos φ cos θ
2 2
= 0.
1 − h(r) ∑ ( )
3
= 2
(ϵaij xj dxi ) g −1 Ta g
r a=1
1 1 − h(r) ∑ ( )
3
=− i 2
(ϵaij xj dxi ) g −1 σa g
2 r a=1
1 1 − h(r) [ 3 2 ( )
=− i (x dx − x2 dx3 ) g −1 σ1 g
2 r2
+ (x1 dx3 − x3 dx1 )(g −1 σ2 g)
]
+ (x2 dx1 − x1 dx2 )(g −1 σ3 g) .
[ ( 1) ( 2)
1 1 − h(r) x x
− i 2
(x 3
dx 2
− x2
dx 3
) + (x 1
dx 3
− x3
dx 1
)
2 r r r
( 3 )]
x
+(x2 dx1 − x1 dx2 )
r
[
1 1 − h(r) 1 3 2
=− i x x dx − x1 x2 dx3 + x1 x2 dx3 − x2 x3 dx1
2 r3
]
+ x2 x3 dx1 − x1 x3 dx2
= 0.
Thus, the (1,1) entry of Ag is the same as the (1,1) entry of g −1 dg (and so,
in particular, ) not depend on the gauge potential A).
( α βdoes
Let g = −β̄ α with α real and depending only on φ and β depending only
( α −β )
on φ and θ. Then g −1 = β̄ α and
αφ dφ βφ dφ + βθ dθ
dg =
−β̄φ dφ − β̄θ dθ αφ dφ
so
α −β αφ dφ βφ dφ + βθ dθ
g −1
dg =
β̄ α −β̄φ dφ − β̄θ dθ αφ dφ
128 2. Physical Motivation
( )
1 (−(1 − cos φ)dθ)i
=−
2
which we write as
( )
1 − 12 (2)i (1 − cos φ)dθ
Ag = − . (2.5.43)
2
A similar calculation for the (1,2) entry gives (Ag )1 and (Ag )2 as shown
below.
(ϕg )1 = (ϕg )2 = 0
f (r) 1
(ϕg )3 = = coth r −
r r
Now for the good part. We have already observed that (A, ϕ) is a globally
defined, smooth field configuration on all of R3 . As r −→ ∞ (i.e., as seen
from a distance), the Higgs field approaches the constant value T3 (because
f (r)
−→ 1), whereas (Ag )1 and (Ag )2 approach 0 (because h(r) −→ 0),
r
while (Ag )3 , which does not depend on r, remains fixed at −(1 − cos φ)dθ. As
seen from infinity the potential function A in this gauge assumes the form
1 − 2 (2)i (1 − cos φ)dθ
1
0
−
2 0 1
(2)i (1 − cos φ)dθ
2
dA ϕ = dϕ + [A, ϕ]
( ) [ ]
f (r) f (r)
=d T3 + Aa Ta , T3
r r
( )
f (r) f (r) a
=d T3 + A [Ta , T3 ]
r r
( )
f (r) f (r) [ 1 ]
=d T3 + A [T1 , T3 ] + A2 [T2 , T3 ] + A3 [T3 , T3 ]
r r
( )
f (r) f (r)
=d T3 + [−A1 T2 + A2 T1 ]
r r
( )
f (r) 2 ∂ f (r)
= (A T1 − A1 T2 ) + dr T3
r ∂r r
( ) ( )
1 1
= coth r − (A2 T1 − A1 T2 ) + − csch 2
r dr T3
r r2
130 2. Physical Motivation
( )
A 1 [
d ϕ = coth r − (r csch r) (cos θ dφ − sin θ sin φ dθ)T1
r
]
+ (sin θ dφ + cos θ sin φ dθ)T2 (2.5.45)
( )
1
+ − csch r dr T3
2
r2
so
∫
1 ( )
A(A, ϕ) = ∥ F ∥2 + ∥ dA ϕ∥2 dx1 ∧ dx2 ∧ dx3
2 R3
∫ ⟨ ⟩
=± F , ∗ dA ϕ dx1 ∧ dx2 ∧ dx3
R3
∫
( )
=∓ 2 trace F ∧ ∗∗dA ϕ
R3
∫
( )
=∓ 2 trace F ∧ dA ϕ
R3
which we now write as
∫
( )
A(A, ϕ) = ∓ Tr F ∧ dA ϕ (monopole), (2.5.46)
R3
where Tr = 2 trace.
∫ Computing this integral
∫ for the t’Hooft-Polyakov monopole
(by writing it as R3 Tr(∗ dA ϕ∧dA ϕ) = − R3 ∥ dA ϕ ∥2 dx1 ∧dx2 ∧dx3 and using
(2.5.45)) gives a value of 4π. For any configuration (A, ϕ) ∈ C satisfying the
monopole equations (2.5.14) we define the monopole number N (A, ϕ) by
∫
1
N (A, ϕ) = Tr(F ∧ dA ϕ). (2.5.47)
4π R3
There are alternative ways of computing N (A, ϕ) (several of which are dis-
cussed below) that make it clear that N (A, ϕ) is actually an integer. Indeed,
such an integer-valued monopole number∫can be defined in a much more gen-
A
R3 Tr(F ∧ d ϕ) is well-defined and
1
eral context. In [JT] it is shown that 4π
integer-valued for any (A, ϕ) ∈ C that is a critical point for the action A given
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 131
by (2.5.10). Then [Groi2] shows that it is not even necessary to assume (A, ϕ)
is a critical point. More precisely, if
∫
1 ( )
A(A, ϕ) = ∥ F ∥2 + ∥ dA ϕ ∥2 dx1 ∧ dx2 ∧ dx3
2 R3
and { }
C= ( A, ϕ) : A ( A, ϕ) < ∞, lim sup 1 − ∥ϕ∥ = 0 ,
R−→∞ | x |≥R
= Tr(F ∧ dA ϕ ),
where the last equality follows from the fact that dA ϕ ∧ F and F ∧ dA ϕ differ
by a bracket, which has trace zero. This proves (2.5.48) and Stokes’ Theorem
(Section 4.7) gives
∫ ( ) ∫ ∫
A
( )
Tr F ∧ d ϕ = d Tr( ϕF ) = Tr( ϕF ).
| x |≤R | x |≤R | x |=R
132 2. Physical Motivation
Writing SR2
for the set of points in R3 with | x | = R we obtain
∫ ∫
1 A 1
N ( A, ϕ) = Tr ( F ∧ d ϕ) = lim Tr ( F ∧ dA ϕ)
4π R3 R−→∞ 4π | x |≤R
∫
1
N ( A, ϕ) = lim Tr (ϕF ). (2.5.49)
R−→∞ 4π 2
SR
and, if R > R0 ,
ϕ̂R = ϕ̂|SR2 .
2
These are smooth maps on their domains and they map into Ssu(2) (the unit
∼
2-sphere in su(2) = R ). One can show (see [JT] and [Groi2]) that ϕ can be
3
The reason for preferring the maps ϕ̂ and ϕ̂R can be seen as follows: Each
ϕ̂R can be regarded as a map from S 2 to S 2 and so determines an element
[ϕ̂R ] of the homotopy group π2 (S 2 ). Since ϕ̂ is smooth for R > R0 , ϕ̂R varies
smoothly with R > R0 so this homotopy class is independent of R > R0
and we will denote it simply [ϕ̂]. We claim that [ϕ̂] is also gauge invariant,
i.e., that if g : R3 −→ SU (2) is a gauge transformation and ϕg = g −1 ϕg,
then, on | x | > R0 , ∥ ϕg ∥−1 ϕg is well-defined and homotopic to ∥ ϕ ∥−1 ϕ. It is
well-defined because ∥ ϕg ∥ = ∥ g −1 ϕg ∥ = ∥ ϕ ∥ which is nonzero on | x | > R0 .
On the other hand, since R3 is contractible, g is homotopic to the map that
sends all of R3 the identity e in SU (2) (Exercise 2.3.6, [N4]). Thus, on | x | >
R0 , ∥ ϕg ∥−1 ϕg = ∥ ϕ ∥−1 (g −1 ϕg) is homotopic to ∥ ϕ ∥−1 (e−1 ϕe) = ∥ ϕ ∥−1 ϕ
so [ϕg ] = [ϕ] as required.
Each map ϕ̂R can be regarded as a map from S 2 to S 2 and therefore has
a Brouwer degree deg(ϕ̂R ) (see Section 5.7 or Section 3.4 of [N4]). Since the
various maps ϕ̂R , R > R0 , determine the same homotopy class in π2 (S 2 ),
they have the same degree. Remarkably, this degree actually coincides with
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 133
(see [JT] and [Groi2]). Notice that deg(∥ ϕ ∥−1 ϕ|SR2 ) depends only on ϕ
and, indeed, only on its asymptotic behavior. Even more, it depends only
on the “homotopy type of its asymptotic behavior” (if you get my drift).
The monopole number distinguishes “homotopy classes” of Higgs fields. These
classes are stable in the sense that a continuous perturbation of the field can-
not change the class (physicists would say that an infinite potential barrier
separates fields with different monopole numbers). Mathematically, there is
a natural topology on the configuration space C with path components la-
beled by the integers and such that two configurations lie in the same path
component if and only if they have the same monopole number (see [Groi2]).
We point out that there is an explicit integral formula for calculating the
degrees (monopole numbers) in (2.5.51) that is sometimes more manageable
than those in earlier formulas:
∫ ( )
1
N (A, ϕ) = − Tr ϕ̂dϕ̂ ∧ dϕ̂ (R > R0 ). (2.5.52)
4π SR2
Thus,
dϕ̂ ∧ dϕ̂
( )( )
1 dx3 i dx2 + dx1 i dx3 i dx2 + dx1 i
=
4R 2 −dx2 + dx1 i −dx3 i −dx2 + dx1 i −dx3 i
( )
1 dx1 ∧ dx2 i −dx1 ∧ dx3 + dx2 ∧ dx3 i
=−
2R 2 dx1 ∧ dx3 + dx2 ∧ dx3 i −dx1 ∧ dx2 i
and so
( )
1 x3 i x2 + x1 i
ϕ̂ dϕ̂ ∧ dϕ̂ =
4R 3 −x2 + x1 i −x3 i
( )
dx1 ∧ dx2 i −dx1 ∧ dx3 + dx2 ∧ dx3 i
× .
dx1 ∧ dx3 + dx2 ∧ dx3 i −dx1 ∧ dx2 i
134 2. Physical Motivation
Thus,
( ) ( )
Tr ϕ̂dϕ̂ ∧ dϕ̂ = 2 trace ϕ̂dϕ̂ ∧ dϕ̂
1 ( 1 2 )
=− 3
x dx ∧ dx3 − x2 dx1 ∧ dx3 + x3 dx1 ∧ dx2 .
R
2
We will learn how to integrate such a 2-form over SR in Chapter 4 (indeed,
we will find that the restriction of x dx ∧ dx − x dx1 ∧ dx3 + x3 dx1 ∧ dx2 to
1 2 3 2
S 2 is just the standard volume (i.e., area) form on S 2 ). Once the machinery
is all in hand we will find that one can calculate such things by simply doing
what comes natural. In this case, one introduces spherical coordinates
x1 = R sin φ cos θ
x2 = R sin φ sin θ
x3 = R cos φ
= R3 sin φ cos θ(cos φ sin θdφ + sin φ cos θdθ) ∧ (− sin φdφ)
= −4π.
2.5. SU (2)-Yang-Mills-Higgs Theory on Rn 135
ϕ̂R = ϕ̂ 2
SR
2
: SR −→ Ssu(2)
2
.
Let ΦR = ΦSR2 ×SU (2) and Φ̂R = ∥ΦR ∥−1 ΦR . Both are equivariant and Φ̂R
2 2
takes values in Ssu(2) . Furthermore, ϕ̂R is the pullback to SR by the standard
cross-section of Φ̂R . Thus, ϕ̂R is the standard gauge representation for a Higgs
2 2
field on the trivial SU (2)-bundle over SR with values in Ssu(2) .
Now, select some ϕ0 ∈ Ssu(2) (a “ground state” for the “virtual potential”;
2
see pages 132–133). The isotropy subgroup of ϕ0 (with respect to the adjoint
action of SU (2) on su(2)) is a copy of U (1) in SU (2) (pages 132–133). One
−1 2
can show that Φ̂R (ϕ0 ) is a submanifold of SR × SU (2) (because Φ̂R is a
−1
submersion at each point of Φ̂R (ϕ0 )) and, furthermore
−1
(i) for each x ∈ SR
2
, P −1 (x) ∩ Φ̂R (ϕ0 ) ̸= ∅, and
136 2. Physical Motivation
−1
(ii) for p ∈ Φ̂R (ϕ0 ) and g ∈ SU (2),
−1
p · g ∈ Φ̂R (ϕ0 ) iff g ∈ U (1) (isotropy subgroup of ϕ0 ).
2 −1
is a principal U (l)-bundle over SR (where the action of U (1) on Φ̂R (ϕ0 )
2
is just the original SU (2)-action on SR × SU (2), but with p restricted to
−1
Φ̂R (ϕ0 ) and g restricted to U (1) ⊆ SU (2)). This U (1)-bundle over SR
2
is
P
called a reduction of the structure group of SU (2) ,→ SR 2
× SU (2) −→ SR 2
to
U (1). Recall that principal U (1)-bundles over spheres are characterized up to
equivalence by their 1st Chern number (see pages 63–64 and 68 of Section 2.2).
The result of interest to us is the following: The 1st Chern number of
−1
−1 P Φ̂R (ϕ0 )
U (1) ,→ Φ̂R (ϕ0 ) −−−−−−−→ SR
2
⟨A, B⟩ = −2trace(AB)
(and γ is “tensorial of type µ”). Computing the 1st Chern number of this
bundle from ω 0 gives the expression (2.5.52) for N (A, ϕ).
2.6 Epilogue
We hope by now to have satisfied the curiosity of those who may have won-
dered how such apparently abstruse mathematical notions as spinor structures
and characteristic classes might arise in the study of the world around us. We
will have one more serious encounter with this in the Appendix, but it is time
now to put aside the informal, heuristic, discussions that have characterized
this chapter and deal honestly with these notions for their own sake. The
remainder of the book is intended to do just that. Certainly, one need not
demand any physical motivation to study and appreciate the rather beautiful
mathematics to follow. Nevertheless, it is pro-foundly satisfying that the phys-
ical motivation exists and, as the concepts are made precise and the theorems
are rigorously proved, we recommend a periodic dip in the murkier waters
of physics (the journal Communications in Mathematical Physics is fine for
browsing). It lends perspective.
3
Frame Bundles and Spacetimes
and relatively compact for i ≤ n and W̄i ⊆ Wi+1 for i ≤ n − 1. Since W̄n
is compact it is contained in a finite union of the Bi . Let in be the least
positive integer greater than or equal to n such that W̄n ⊆ B1 ∪ · · · ∪ Bin .
Set Wn+1 = B1 ∪ · · · ∪ Bin . Then Wn+1 is open, W̄n ⊆ Wn+1 and W̄n+1 =
B1 ∪ · · · ∪ Bin = B̄1 ∪ · · · ∪ B̄in is compact so the induction is complete and
we have a sequence {W1 , W2 , . . .} of relatively compact open sets and with
W̄i ⊆ Wi+1 for each i = 1, ∪2, . . . . Since Wi contains the union of the first i
∞
elements of {B1 , B2 , . . .}, i=1 Wi = X and {W1 , W2 , . . .} is a cover of X.
Exercise 3.1.2 Show that each of the sets W̄2 , W̄3 − W2 , W̄4 − W3 , . . . ,
W̄i − Wi−1 , . . . is compact, each of the sets W3 , W4 − W̄1 , W5 − W̄2 , . . .,
Wi+1 − W̄i−2 , . . . is open, and
W̄2 ⊆ W3
W̄3 − W2 ⊆ W4 − W̄1
W̄4 − W3 ⊆ W5 − W̄2
..
.
W̄i − Wi−1 ⊆ Wi+1 − W̄i−2
..
.
With this auxiliary cover in hand we can now prove the lemma. Let U =
{Uα : α ∈ A} be an arbitrary open cover of X. Define U2 = {Uα ∩W3 : α ∈ A}
and, for i ≥ 3, Ui = { Uα ∩ (Wi+1 − W̄i−2 ) : α ∈ A}. Then U2 covers W̄2
and, for i ≥ 3, Ui covers W̄i − Wi−1 . Let V2 be a finite collection of ele-
ments of U2 that cover W̄2 and, for i ≥ 3, let Vi be a∪finite collection of
∞
elements of Ui that cover W̄i − Wi−1 . Finally, let V = i=2 Vi . Then V is
a countable family of open sets in X. Each element of V is contained in
some Uα as well as in some W̄j so that V refines U and the elements of V
have compact closure. To see that V is a cover, let x be an arbitrary el-
ement of X. Select the least integer i ≥ 1 such that x ∈ Wi . If i = 1,
or 2, then x ∈ W̄2 and therefore in some element of V2 . If i ≥ 3, then
x ∈ W̄i − Wi−1 so x is in some element of Vi . This also implies that V is
locally finite since any x ∈ X is in one of the compact sets W̄2 , W̄3 − W2 , . . .
and the corresponding open set W3 , W4 − W̄1 , . . . is an open neighbor-
hood of x which, by construction, intersects only finitely many elements
of V.
locally finite, open refinement V = {Vk }k=1,2,... of U ′ with each V̄k compact.
Thus, V is also a refinement of U and, moreover, each V̄k is contained in
some element of U. By Exercise 3.1.1, {V̄k }k=1,2... is also locally finite so, to
complete the proof, it will suffice to find a partition of unity {ϕk }k=1,2... with
supp ϕk = V̄k for each k = 1, 2, . . . .
For each k = 1, 2, . . . select α(k) ∈ A such that V̄k ⊆ Uα(k) . Then Uα(k) − Vk
is a nonempty, proper closed subset of Uα(k) so, by Lemma 3.1.2, there is a non-
negative smooth function fk′ on Uα(k) with (fk′ )−1 (0) = Uα(k) − Vk . From this
we construct a non-negative smooth function fk on X with fk−1 (0) = X − Vk
as follows: For each p ∈ V̄k select a non-negative bump function gp which
is 1 on a neighborhood Up of p in Uα(k) and 0 on X − Uα(k) . Cover V̄k by
finitely many of these neighborhoods Up1 , . . . , Upj and let g = gp1 + · · · + gpj .
Then g is C ∞ on X, nonzero on V̄k and 0 on X − Uα(k) . Define fk on X
by
fk′ (x)g(x), x ∈ Uα(k)
fk (x) = .
0, x ∈ X − Uα(k)
for each x ∈ X. Moreover, since every x ∈ X is in some Vk , f (x) > 0 for each
x ∈ X. Thus, for each k = 1, 2, . . . , the function ϕk : X −→ R defined
by
fk (x)
ϕk (x) =
f (x)
is non-negative, C ∞ , has supp ϕk = supp fk
∑ = V̄k and satisfies
∞
k=1 ϕk (x) = 1 for each x ∈ X as required.
The functions in a partition of unity are required to have compact support.
On occasion it is convenient to drop this requirement.
Proof: Observe first that if (U, φ) is a chart on X, then the open submanifold
U of X admits a Riemannian metric (e.g., φ∗ ḡ , where ḡ is the standard
metric on φ(U ) ⊆ Rn ). Now, let {(Uα , φα ) : α ∈ A} be an atlas for X.
By Theorem 3.1.4 there is a countable partition of unity {ϕk }k=1,2,... on
X subordinate to {Uα : α ∈ A}. For each k = 1, 2, . . . choose α(k) ∈ A
such that supp ϕk ⊆ Uα(k) . Then {Uα(k) }k=1,2,... is a countable subcover of
{Uα : α ∈ A}. On each Uα(k) select a Riemannian metric g k and define g by
∞
∑
g= ϕk g k .
k=1
of any
∑∞one of these coordinate neighborhoods Uα(k) . Then, on U , the sum
g = k=1 ϕk g k defining g is finite, each ϕk is a C ∞ function of x1 , . . . , xn
∂ ∂ ∞ ∂ ∂ ∞
and each g k ( ∂x i , ∂xj ) is C so g ( ∂x i , ∂xj ) is C for i, j = 1, . . . , n. Thus,
g is smooth on U as required.
where it is understood that, if P(p) ̸∈ supp ϕk , then the term ϕk (P(p))ω k (p)
(v) is taken to be zero. Note that (3.1.2) does, indeed, define a G-valued 1-
form on P since, at each p, the sum is finite and G is a real vector space.
as required.
Notice that
∑
h(x) = ϕ0 (x) + ϕp (x) h(x)
p∈U −A
∑
= ϕ0 (x) h (x) + ϕp (x) h (x)
p∈U −A
so ∑
f (x) − h(x) = ϕp (x) (h(p) − h(x)).
p∈U −A
is a Lie group, X and P are smooth manifolds and P as well as the action of
P
G on P are smooth. Then G ,→ P −→ X is a smooth principal bundle (i.e.,
has smooth local trivializations). One need only apply the result from [St] to
each trivialization.
We conclude with two more items of this same sort. Given a smooth man-
ifold X, a Lie group G, an open cover {Vj }j∈J of X and a family {gji } of
smooth maps from nonempty intersections Vi ∩ Vj into G satisfying the co-
cycle condition, it is a simple exercise to trace through the proof of the (C 0 )
Reconstruction Theorem (Theorem 4.3.4, [N4]) and show that the resulting
bundle admits a natural smooth structure (if such an exercise does not appeal
to you, see Proposition 5.2, Chapter I, of [KN1]). One thereby obtains the ver-
sion of the Reconstruction Theorem recorded on page 34. Similarly, the proof
of the (C 0 ) Classification Theorem for principal G-bundles over S n (Theo-
rem 4.4.3, [N4]) shows that the equivalence class of such a bundle is uniquely
determined by the homotopy type of its characteristic map T = g12 | S n−1
from the equator S n−1 ⊆ S n into G. If G is a Lie group one can choose a
smooth map homotopic to T and use this as a characteristic map to build
a G-bundle over S n (Lemma 4.4.1, [N4]). This latter bundle is smooth (by
the smooth Reconstruction Theorem) and equivalent to the original bundle
so every equivalence class contains a smooth bundle. In this way one arrives
at the smooth version of the Classification Theorem, as stated on page 34.
b̂ j = b i g i j , j = 1, . . . , n. (3.3.1)
Formally, we may write
g 11 ··· g 1n
.. ..
(b̂ 1 · · · b̂ n ) = (b 1 · · · b n ) . . (3.3.2)
g n1 ··· g nn
p · g = p ◦ g : Rn −→ Tx (X) (3.3.5)
g e e
since ( )
p ◦ g (ej ) = p(ei g ij ) = p(ei )g ij = b i g ij = b̂ j
e e e e
for j = 1, . . . , n.
Our objective now is to provide L(X) with a topology and manifold struc-
ture in such a way that, with the action σ described above,
P
GL(n, R) ,→ L(X) −→
L
X
is a smooth principal GL(n, R)-bundle over X, called the (linear) frame
bundle of X. Toward this end we let (U, φ) be any chart on X with coordinate
functions x1 , . . . , xn . Define φ̃: PL−1 (U ) −→ φ(U ) × GL(n, R) as follows: Let
150 3. Frame Bundles and Spacetimes
∂
bj = Ai j (p).
∂xi x
The matrix ( )
A(p) = Ai j (p)
on L(X) by declaring that a subset U of L(X) is open if and only if, for each
chart (U, φ) on X, φ̃(U ∩ PL−1 (U )) is open in φ̃(PL−1 (U )) = φ(U ) × GL(n, R).
Exercise 3.3.1 Show that the collection of all such subsets U of L(X) does,
indeed, define a topology for L(X) and that, if (V, ψ) is any chart for X, then
PL−1 (V ) is open in L(X).
Next we consider two charts (U, φ) and (V, ψ) for X with coordinate func-
tions x1 , . . . , xn and y 1 , . . . , y n , respectively, and with U ∩ V ̸= ∅. Then φ̃ and
ψ̃ are both defined on PL−1 (U ∩ V ), which is open in L(X). We compute
∂
bj = Ai j , j = 1, . . . , n.
∂xi x
∂
bj = B ij , j = 1, . . . , n.
∂y i x
But
( ) ( )
∂ ∂ ∂y i ∂ ∂y i
bj = Akj = (x) Akj = (x)Akj
∂xk x ∂y i x ∂xk ∂y i
x ∂xk
so
∂y i
B ij = (x) Akj , i, j = 1, . . . , n.
∂xk
3.3. Frame Bundles 151
Thus,
( )
ψ̃ ◦ φ̃−1 ((x1 , . . . , xn ), (Ai j ))
(3.3.7)
( ( i ))
∂y
= (ψ ◦ φ−1 )(x1 , . . . , xn ), (x) A k
j ,
∂xk
which, in particular, is C ∞ . Reversing the roles of ψ̃ and φ̃ we find that ψ̃◦ φ̃−1
and φ̃ ◦ ψ̃ −1 are inverse diffeomorphisms and we will use this fact to provide
L(X) with a differentiable structure.
First we show that, for any chart (U, φ) on X, the map φ̃ : PL−1 (U ) −→
φ(U ) × GL(n, R) is a homeomorphism (so that, in particular, L(X) is locally
Euclidean). We need only show that φ̃ is continuous and an open map. For the
latter, we let W be an open set in PL−1 (U ). Since PL−1 (U ) is open in L(X), W
is open in L(X) so φ̃(W ∩ PL−1 (U )) = φ̃(W ) is open in φ(U ) × GL(n, R),
as required. To prove continuity we let Z be open in φ(U ) × GL(n, R). To
show that φ̃−1 (Z) is open in PL−1 (U ) we let (V, ψ) be an arbitrary chart for
X with U ∩ V ̸= ∅. We must show that ψ̃(φ̃−1 (Z) ∩ PL−1 (V )) is open in
ψ(V ) × GL(n, R). But
( )
ψ̃(φ̃−1 (Z) ∩ PL−1 (V )) = ψ̃ φ̃−1 (Z) ∩ PL−1 (U ) ∩ PL−1 (V )
( )
= ψ̃ φ̃−1 (Z) ∩ PL−1 (U ∩ V )
( )
= ψ̃ φ̃−1 (Z) ∩ φ̃−1 (φ(U ∩ V ) × GL(n, R))
( )
= ψ̃ φ̃−1 (Z ∩ (φ(U ∩ V ) × GL(n, R)))
( )
= (ψ̃ ◦ φ̃−1 ) Z ∩ (φ(U ∩ V ) × GL(n, R))
(U, φ) on X we define
Φ : PL−1 (U ) −→ U × GL(n, R)
∂
bj = Ai j (p), j = 1, . . . , n.
∂xi x
But then
∂
b̂ j = b k g k j = Ai k (p)g k j
∂xi x
implies
Ai j (p · g) = Ai k (p)g k j , i, j = 1, . . . , n.
Since the right-hand side of this last equality is the (i, j)-entry in the matrix
product A(p)g we have A(p · g) = A(p)g, as required.
Notice that, if Φ is the trivialization arising from (U, φ), with coordinate
functions x1 , . . . , xn and Ψ is the trivialization arising from (V, ψ) with coor-
dinate functions y 1 , . . . , y n and if U ∩ V ̸= ∅, then it follows from (3.3.7) that
the transition function gV U : U ∩ V −→ GL(n, R) is just the Jacobian of the
coordinate transformation, i.e.,
( i )
∂y
gV U (x) = (x)
∂xk
gn 1 ··· gn n
154 3. Frame Bundles and Spacetimes
and g −1 · v is given by
−1
v̂ 1 g1 1 ··· g1 n v1
. . .. ..
. = .
. . . .
v̂ n gn 1 ··· gn n vn
g1 1 ··· gn 1 v1
. .. ..
= .
. . . .
g1 n ··· gn n vn
is just the set of all possible descriptions of some fixed tangent vector at x.
Note that, despite the ultramodern attire, this is just a dressed-up version of
the “old-fashioned” view of a vector as a collection of n-tuples, one for each
basis, related by the transformation law ρ.
A vector field on an open subset U of X can now be identified with a
local cross-section V : U −→ T (X) of the tangent bundle. Equivalently (Sec-
tion 6.8, [N4]), one can identify V with an Rn -valued map on PL−1 (U ) that
is equivariant, i.e., satisfies V (p · g) = g −1 V (p) for each p ∈ PL−1 (U ) and
g ∈ GL(n, R).
Similarly, if one defines a representation ρ : GL(n, R ) −→ GL(Rn ) by
ρ(g)(θ) = g · θ = (g T )−1 θ, where θ ∈ Rn is written as a column matrix and
(g T )−1 θ denotes matrix multiplication, then the associated vector bundle is
called the cotangent bundle and denoted T ∗ (X). A 1-form can then be iden-
tified with either a cross-section of T ∗ (X) or an Rn -valued map on L(X) that
is equivariant with respect to this representation (θ(p · g) = g T θ(p)). Tensor
bundles (and their cross-sections, or equivariant maps, called tensor fields)
arise in exactly the same way by making other choices for the representation ρ.
If an n-dimensional manifold X has a Riemannian metric g defined on
it, then a frame p = (b 1 , . . . , b n ) at x ∈ X is orthonormal if g (b i , b j ) =
δij , i, j = 1, . . . , n. Such frames are related by elements of the orthogonal group
O(n) ⊆ GL(n, R) and we wish to build an “orthonormal frame bundle” with
P
group O(n) analogous to the linear frame bundle GL(n, R) ,→ L(X) −−→ L
X.
However, we will need the construction in the indefinite case as well so we
begin by generalizing what we know about the positive definite case. Most of
3.3. Frame Bundles 155
the proofs are virtually identical and so will be left to the reader in a sequence
of exercises.
On Rn we will denote by ⟨ , ⟩k , 0 ≤ k ≤ n, the standard inner product of
index n − k. Thus, if {e1 , . . . , en } is the standard basis for Rn , x = xi ei and
y = y j ej , then
⟨x, y⟩k = x1 y 1 + · · · + xk y k − xk+1 y k+1 − · · · − xn y n = ηij xi y j ,
where
( )
idk×k 0
η = (ηij ) = .
0 −id(n−k)×(n−k)
Exercise 3.3.6 Show that O(k, n − k) is, indeed, a group under matrix mul-
tiplication and that det A = ±1 for every A ∈ O(k, n − k).
The subgroup
{ }
SO(k, n − k) = A ∈ O(k, n − k) : det A = 1
is called the special semi-orthogonal group. Of course, when k = n we
write O(n, 0) = O(n) and SO(n, 0) = SO(n). We will have more to say
later about O(1, 3), called the general Lorentz group and denoted L, and
SO(1, 3), called the proper Lorentz group and denoted L+ .
Exercise 3.3.7 Show that, for any θ ∈ R, the matrix
cosh θ 0 0 − sinh θ
0 1 0 0
L(θ) =
0 0 1 0
− sinh θ 0 0 cosh θ
is in L+ .
156 3. Frame Bundles and Spacetimes
O(k, n − k) = f −1 (η).
From Exercise 3.3.8 (and Corollary 5.6.7, [N4]) we conclude that O(k, n−k)
is a submanifold of GL(n, R) ∼ = Rn of dimension n2 − 12 n(n + 1) =
2
a Lie group.
Remark: In the positive definite case (k = n), the orthogonal group O(n)
is compact (because the rows of any A ∈ O(n) must form a Euclidean or-
thonormal basis for Rn so O(n) ⊆ Rn is bounded). This is not the case
2
when 0 < k < n, e.g., the elements of O(1, 3) described in Exercise 3.3.7 form
an unbounded set.
Since the determinant function on GL(n, R) is continuous and
SO(k, n − k) = O(k, n − k) ∩ det−1 (0, ∞), the special semi-orthogonal group
is an open submanifold of O(k, n − k) and therefore is also a Lie group.
Since SO(k, n − k) is an open submanifold of O(k, n − k), the Lie algebras
so(k, n − k) and o(k, n − k) are the same. One determines this Lie algebra in
precisely the same way as for O(n) and SO(n).
Exercise 3.3.9 Mimic the arguments on pages 279–280 of [N4] to show that
so(k, n − k) = o(k, n − k) = {A ∈ GL(n, R) : AT = −ηAη}.
Now, a metric g on an n-manifold X is said to be semi-Riemannian of
index n − k if, at each x ∈ X, the inner product g x has index n − k. Then a
frame p = (b 1 , . . . , b n ) at x is orthonormal if
1, i = j = 1, . . . , k
g x (b i , b j ) = ηij = −1, i = j = k + 1, . . . , n .
0, i ̸= j
We denote by F (X)x the set of all orthonormal frames at x and let F (X) =
∪
x∈X F (X)x . For each p ∈ F (X)x ⊆ F (X) we let PF (p) = x and thereby
obtain a surjective map
PF : F (X) −→ X.
Next we define a right action σ : F (X)×O(k, n−k) −→ F (X) of O(k, n−k) on
F (X) as follows: For each (p, g) ∈ F (X)×O(k, n−k), with p = (b 1 , . . . , b n ) ∈
F (X)x ⊆ F (X) and g = (g i j ) ∈ O(k, n − k), we let σ(p, g) = p · g ∈ F (X)x
be the frame (b̂ 1 , . . . , b̂ n ) at x, where
b̂ j = b i g ij , j = 1, . . . , n.
3.3. Frame Bundles 157
Notice that this frame is, indeed, orthonormal by Exercise 3.3.4 and that
σ is a right action for precisely the same reason that the analogous map on
L(X)×GL(n, R) is a right action (see (3.3.2)). By definition, PF (p·g) = PF (p)
for all (p, g) ∈ F (X)×O(k, n−k). Just as for L(X) it will sometimes be conve-
nient to identify an orthonormal frame p with the corresponding isomorphism
p : Rk,n−k −→ Tx (X).
e
We wish to show that
P
O(k, n − k) ,→ F (X) −−→
F
X
b j = E i (x)Ai j (p).
Exercise 3.3.11 Show that the collection of all such subsets U of F (X) does,
indeed, define a topology on F (X) and that, if (V, ψ) is any chart for X, then
PF−1 (V ) is open in F (X).
Next let (U, φ) and (V, ψ) be two charts on X with U ∩ V ̸= ∅ and with
orthonormal frame fields {E 1 , . . . , E n } and {F 1 , . . . , F n }, respectively. For
each x ∈ U ∩ V we write
is given by
( )( )
ψ̃ ◦ φ̃−1
(x1 , . . . , xn ), (Ai j )
( ( ))
= (ψ ◦ φ−1 )(x1 , . . . , xn ), Λki (x)Akj ,
for any semi-Riemannian manifold X and we now have every confidence that
the reader can take the next step without assistance.
Exercise 3.3.17 Let X be an oriented semi-Riemannian manifold of index
n − k. Construct, in detail, the oriented, orthonormal frame bundle
PF+
SO(k, n − k) ,→ F+ (X) −→ X.
Being a Lie group S 3 has a trivial linear frame bundle. It is instructive, and
will be useful somewhat later, to explicitly construct a cross-section of the
oriented, orthonormal frame bundle, thereby showing that it is trivial as well.
Exercise 3.3.18 Show that it will suffice to define smooth vector fields
V 1 , V 2 and V 3 on S 3 with the property that, for each p ∈ S 3 , {V 1 (p),
V 2 (p), V 3 (p)} is an oriented, orthonormal basis for Tp (S 3 ).
The procedure for constructing the vector fields described in Exercise 3.3.18
is quite simple. Regard S 3 ⊆ R4 = H as the unit quaternions. For each p ∈ S 3 ,
the tangent space Tp (S 3 ) can be identified with a subspace of Tp (R4 ) and this,
in turn, is canonically identified with R4 itself. Viewed in this way, Tp (S 3 ) is
just the subspace of Tp (R4 ) = R4 consisting of those v = (v 0 , v 1 , v 2 , v 3 ) with
⟨v, p⟩ = 0, i.e., withv 0 p0 + v 1 p1 + v 2 p2 + v 3 p3 = 0. Now, select a point, say
p0 = (1, 0, 0, 0), in S 3 and let
is positive. One can compute this directly or argue indirectly as follows: Since
the rows of (3.3.10) are orthogonal unit vectors in R4 the matrix itself is
an orthogonal matrix and so has determinant ±1 at each p ∈ S 3 . But the
determinant function is continuous and S 3 is connected so the determinant of
(3.3.10) is either 1 for all p ∈ S 3 or −1 for all p ∈ S 3 . Since this determinant
is obviously 1 when p = (1, 0, 0, 0) ∈ S 3 the result follows.
We will have occasion somewhat later to introduce yet one more frame
bundle (the “oriented, time oriented, orthonormal frame bundle” of a “space-
time” manifold). For the present we will conclude this discussion with a
few remarks on some topics we will not pursue in any depth here. Accord-
ing to Theorem 3.1.7, any principal bundle has connections defined on it.
and this is true, in particular, for the frame bundles we have constructed.
A connection on the linear frame bundle GL(n, R) ,→ L(X) −→ X is called
a linear connection on X and these are of fundamental importance in the
study of the geometry of X (see Chapter III of [KN1]). If X admits a Rie-
mannian or semi-Riemannian metric, then there is, among these linear con-
nections, a distinguished one called the Levi-Civita connection (or Rie-
mannian connection) that is adapted to the metric structure. The study
of this connection is the vast and beautiful subject of (semi-) Riemannian
geometry. Although this is not our subject here we intend to borrow one of
its results. Moreover, the Levi-Civita connection plays a role in defining the
Seiberg-Witten equations which we will sketch in the Appendix so a brief tour
of the definition is probably in order (for details one can consult [Bl], [O’N],
or [KN1]). We consider a smooth n-manifold X and its linear frame bundle
GL(n, R) ,→ L(X) −→ X. A connection ω on L(X) is a gl(n, R)-valued 1-
form on L(X) which we describe as follows. Let Eij be the n × n matrix for
which the entry in the ith row and j th column is 1 and all other entries are 0.
Then ω can be written as ω = ω ij Eij , where each ω ij is a real-valued 1-form on
L(X). Similarly, the curvature Ω = dω + 12 [ω, ω] can be written Ω = Ω ij Eij ,
where each Ω ij is a real-valued 2-form on L(X). Thus,
Ω ij = dω ij + ω ik ∧ ω kj .
3.3. Frame Bundles 161
i
for some smooth functions Γjk on U (called the Christoffel symbols for ω
in the coordinate neighborhood U ). Similarly,
( )
∗ ∗ 1 i
Ω U = (SU ) Ω = ((SU ) Ω ij )Eij = (R dxk ∧ dxl )Eij
2 jkl
i
for some smooth functions Rjkl on U . Unraveling the definitions gives
i
Rjkl = ∂k Γlji − ∂l Γkj
i
+ Γljm Γkm
i
− Γkj
m i
Γlm
for i, j, k, l = 1, . . . , n.
Now suppose that X has a metric g of index n − k and consider the
orthonormal frame bundle O(k, n−k) ,→ F (X) −→ X. If X is also oriented we
have an oriented, orthonormal frame bundle SO(k, n − k) ,→ F+ (X) −→ X as
well. It is not difficult to see that any connection on F (X) or F+ (X) extends
uniquely to a connection on L(X) (page 158 of [KN1]), but it is not true
that every connection on L(X) restricts to a connection on F (X) or F+ (X)
because a connection on L(X) takes values in gl(n, R) and not necessarily in
o(k, n − k) = so(k, n − k); those that do are called metric connections on
L(X). The so-called Fundamental Theorem of Riemannian Geometry
states that there is a unique metric connection on L(X) that is also symmetric
i i
in the sense that Γjk = Γkj for all i, j, k = 1, . . . , n in any local coordinate
system. This is called the Levi-Civita connection and denoted ω LC (the
same terminology and notation is used for its restriction to either F (X) or
F+ (X)). One can show that ω LC is characterized by the fact that, in any local
coordinate system, the Christoffel symbols are given in terms of the metric
components by
1
i
Γjk = g il (∂k gjl + ∂j gkl − ∂l gjk )
2
for i, j, k, l = 1, . . . , n. All of the usual objects of study in (semi-)
Riemannian geometry are defined in terms of the Levi-Civita connection. For
i
example, Rjkl are the local components of the Riemann curvature tensor,
Rij = Rikj are the components of the Ricci tensor, and R = g ik Rik is the
k
scalar curvature.
162 3. Frame Bundles and Spacetimes
U1
U2
U3
Remark: A much more thorough study of both the mathematics and the
physics is available in [N3].
We will bow to the generally accepted conventions of physics and use
(x0 , x1 , x2 , x3 ), rather than (x1 , x2 , x3 , x4 ), for the standard coordinates in
R1,3 and will drop the subscript 1 on the inner product ⟨ , ⟩1 . Thus, if
{e0 , e1 , e2 , e3 } is the standard basis with x = xα eα and y = y β eβ , then
⟨x, y⟩ = x0 y 0 − x1 y 1 − x2 y 2 − x3 y 3 = ηαβ xα y β
where
1, α=β=0
ηαβ = −1, α = β = 1, 2, 3 .
0, α ̸= β
eβ = Λα β êα , β = 0, 1, 2, 3,
x̂α = Λα β xβ , α = 0, 1, 2, 3, (3.4.1)
so
Λ0 0 ≥ 1 or Λ0 0 ≤ −1. (3.4.3)
Those elements of L for which Λ 0 ≥ 1 are called orthochronous and the
0
set
L+↑ = {Λ ∈ L+ : Λ0 0 ≥ 1}
is called the proper orthochronous Lorentz group. For physical reasons
that we will discuss shortly we will consider only those inertial observers
(orthonormal bases) related to the standard basis by an element of L+↑ . We
call such bases admissible.
If gravitational effects are assumed negligible, then the Minkowski
inner product ⟨ , ⟩ on R1,3 has something physically significant to say about
relationships between events. Consider, for example, two events x0 , x1 ∈ R1,3
and the displacement vector x = x1 − x0 between them. Suppose that
Q(x) = 0. Then, if we write x = (∆xα )eα ,
and the same equation is satisfied in any other orthonormal basis. Conse-
quently, the spatial separation of the two events x0 and x1 is numerically
equal to the distance light would travel during the time lapse between them.
The two events are “connectible by a light ray.” In this case, x is said to be
null, or lightlike. With x0 held fixed the set of all x1 for which x = x1 − x0
is null is called the null cone, or light cone, at x0 because of the formal re-
semblance of (3.4.4) to the equation of a right circular cone in R3 . A straight
line which lies entirely on such a null cone is called a worldline of a photon
and is thought of as the set of all events in the history of some “particle of
light.”
Suppose instead that Q(x) = Q(x1 − x0 ) > 0. In this case we say that x is
timelike and, in any orthonormal basis, we have
Notice that Exercise 3.4.2 does not assert that admissible observers agree on
the temporal order of x0 and x1 if x1 − x0 is spacelike and, indeed, they need
not.
Exercise 3.4.3 Show that if Λ ∈ L is orthochronous, but det Λ = −1 (see
Exercise 3.3.6), then
1 0 0 0
0 1 0 0
Λ
0 0 1 0
0 0 0 −1
is in L+↑ .
Since the matrix in Exercise 3.4.3 simply reverses the orientation of the spatial
axes, our assumption that admissible observers are related by elements of L+↑
essentially amounts to the requirement that no one’s clock runs backwards
and no one uses “left-handed” spatial coordinates. Henceforth, we will refer to
the elements of L+↑ simply as Lorentz transformations.
Exercise 3.4.4 Show that if (Ri j )i,j=1,2,3 is an element of SO(3), then
1 0 0 0
0
R=
0
(Ri j )
0
is in L+↑ . Show also that the collection of all such elements of L+↑ is a subgroup
R of L+↑ (called the rotation subgroup of L+↑ ).
If two observers are related by an element of R, then they differ only in having
spatial coordinate axes that are rotated relative to each other.
Exercise 3.4.5 Let Λ = (Λα β )α,β=0,1,2,3 be an element of L+↑ . Show that the
following are equivalent.
(a) Λ ∈ R.
(b) Λ1 0 = Λ2 0 = Λ3 0 = 0.
(c) Λ0 1 = Λ0 2 = Λ0 3 = 0.
(d) Λ0 0 = 1.
From the point of view of physics, the elements of R are rather dull. On the
other hand, the matrices L(θ), θ ∈ R, described in Exercise 3.3.7 are clearly
in L+↑ (cosh θ ≥ 1) and these are not at all dull. Physically, they correspond
3.4. Minkowski Spacetime 167
to inertial observers whose spatial coordinate axes are parallel and whose
relative motion is along their common x1 -, x̂1 -axis with speed β = tanh θ (see
Section 1.3 of [N3]). These Lorentz transformations L(θ) are called boosts
and, in some sense, contain all of the interesting kinematical information in
L+↑ .
Theorem 3.4.2 Let Λ be an element of L+↑ . Then there exists a real number
θ and two rotations R1 and R2 in R ⊆ L+↑ such that
Λ = R1 L(θ)R2 .
The intuitive content of this result is quite simple. The Lorentz transformation
Λ from the frame of reference of observer O1 to the frame of reference of
observer O2 can be accomplished in three stages: Rotate O1 ’s spatial axes so
that the x1 -axis coincides with the line along which the relative motion takes
place. Boost to a new frame whose spatial axes are parallel to the rotated
axes of O1 and at rest relative to O2 . Finally, rotate these new spatial axes
so that they coincide with those of O2 . For a detailed algebraic proof see
Theorem 1.3.5 of [N3].
Remark: One can use the decomposition in Theorem 3.4.2 to define a defor-
mation retraction of L+↑ onto SO(3) and conclude that these two have the same
homotopy type (Lemma 2.4.9, [N4]). Indeed, one can show more. L+↑ is actu-
ally homeomorphic to SO(3) × R3 (there is a nice proof of this on pages 73–74
of [Bl]). From either of these it follows that π1 (L+↑ ) ∼
= π1 (SO(3)) ∼
= Z2 (Ap-
pendix B of [N3]). We will need this fact only for motivational purposes in Sec-
tion 6.5 and so will not give the details here. However, we make the following
observation. On pages 92–93 we exhibited two smooth loops R1 (t) and R2 (t)
in SO(3) corresponding to a continuous rotation about the x-axis through
2π and 4π, respectively, and showed that they represent the two equivalence
classes in π1 (SO(3)) ∼
= Z2 . The same two curves, now thought of as loops in
R ⊆ L+ as in Exercise 3.4.4 represent the two classes in π1 (L+↑ ) ∼
↑
= Z2 .
We have already described (on pages 101–103) an alternative model of
Minkowski spacetime as the real linear space H of 2 × 2 complex Hermi-
tian matrices with Q given by the determinant (and ⟨ , ⟩ thereby determined
via the Polarization Identity). This view of R1,3 is particularly convenient for
describing the double cover
Spin : SL(2, C) −→ L+↑
of L+↑ . This, the reader may recall, was required to produce a (“2-valued”) rep-
resentation of L+↑ that is the appropriate transformation law (under change
of inertial frame of reference) for the wavefunction of a spin 12 particle
(Section 2.4). We also briefly alluded (on pages 113–115) to the difficulties
involved in generalizing these considerations to the situation in which gravi-
tational fields are not negligible so that R1,3 is no longer an accurate model
168 3. Frame Bundles and Spacetimes
about the earth. The objects inside the elevator (capsule) seem then to con-
stitute an archetypical inertial frame (they satisfy Newton’s First Law). By
establishing spatial and temporal coordinate systems in the usual way our
observer thereby becomes an inertial observer, at least within the spatial and
temporal constraints imposed by his circumstances. Now picture an arbitrary
event. There are any number of vantage points from which the event can be
observed. One is from a freely falling elevator in the immediate spatial and
temporal vicinity of the event and from this vantage point the event receives
inertial coordinates. There is then a local inertial frame near any event in X.
The operative word is “local.” The “spatial and temporal constraints” to
which we alluded in the preceding paragraph arise from the nonuniformity of
any gravitational field in the real world. For example, in an elevator which
falls freely in the earth’s gravitational field, all of the objects inside are pulled
toward the earth’s center so that these objects do experience some slight
relative motion (toward each other). Such motion, of course, goes unnoticed
if the elevator falls neither too far nor too long. Indeed, by restricting our
observer to a sufficiently small region in space and time these effects become
negligible and the observer is indeed inertial. But then, what is “negligible” is
in the eye of the beholder. The availability of more sensitive measuring devices
will require further restrictions on the size of the spacetime region which “looks
like” R1,3 . Turn of the century mathematical terminology expressed this fact
by saying that any point in X has about it an “infinitesimal neighborhood”
which is identical to R1,3 . Today we prefer to say that X is a 4-dimensional
smooth manifold, each tangent space of which has the structure of R1,3 .
A spacetime is a 4-dimensional smooth manifold X with a semi-Riemannian
metric g of index 3 (called a Lorentz metric). Thus, for each x ∈ X there
exists a frame p = (b 0 , b 1 , b 2 , b 3 ) at x such that
1, α=β=0
g x (b α , b β ) = ηαβ = −1, α = β = 1, 2, 3. (3.5.1)
0, α ̸= β
gαβ (x) = g x ( ∂x∂α |x , ∂x∂ β |x ) = ηαβ for each x ∈ R1,3 is tantamount to defining
the Minkowski inner product on R4 . We shall feel free to think of Minkowski
spacetime in whichever of these ways is convenient at the moment.
The most serious obstacle to the general study of spacetime manifolds is
their overwhelming number and diversity. Smooth 4-manifolds are, to say
the least, plentiful and almost all of them admit Lorentz metrics. Indeed, a
Lorentz metric can be defined on any noncompact 4-manifold and a compact
4-manifold admits a Lorentz metric if and only if its Euler characteristic (see
Section 5.7 or Section 3.4 of [N4]) is zero (see [O’N]).
Remark: Compact spacetimes are of no real interest anyway since they al-
ways contain closed timelike curves and these do violence to our most cherished
notions of causality (see [N2]). In effect then, anything that could possibly be
a spacetime, is a spacetime.
There are two ways around this difficulty. One can impose additional restric-
tions on the structure of a spacetime in an attempt to eliminate “unphysical”
behavior. A number of such restrictions are easily formulated (we will intro-
duce one shortly) and the study of the manifolds that satisfy these conditions
has led to some spectacular results (e.g., the Singularity Theorems of Penrose
and Hawking discussed in [HE], [Pen], [O’N] and [N2]). Alternatively, one
can restrict attention to spacetime manifolds that arise in physics by solving
the field equations of general relativity for more or less realistic distributions
of mass/energy. We will describe a few such examples shortly, but our real in-
terest here (and in Chapter 6) is in an essentially global, topological question
about spacetime manifolds. The question (introduced in Section 2.4) is this:
On which spacetime manifolds is it possible to introduce a meaningful notion
of a “spin 12 particle” ?
According to Dirac a spin 12 particle is characterized by the fact that its
wavefunction transforms under a certain (reducible) representation of the dou-
ble cover SL(2, C) of L+↑ . Now, each tangent space Tx (X) to a spacetime
manifold X has an inner product ⟨ , ⟩x of index 3 giving it the structure
of Minkowski spacetime. In particular, its orthonormal bases are related by
elements of L. We are, however, interested only in bases related by elements of
L+↑ . We must therefore choose, consistently over all of X, a family of bases for
each Tx (X) that are related by the Λ ∈ L satisfying det Λ = 1 and Λ00 ≥ 1.
This is possible only if X is orientable and has, moreover, some global notion
of “time orientation” (see Exercise 3.4.2). To define such a notion we proceed
in the manner familiar for smooth surfaces in R3 (which are orientable if and
only if they admit a smooth, nonzero field of normal vectors).
We say that a spacetime X is time orientable if one can define on it a
smooth vector field T which is everywhere timelike (⟨T (x), T (x)⟩x > 0 for
each x ∈ X). X is time oriented if a specific choice of such a vector field has
been made. One then thinks of T as “pointing toward the future” (keeping
in mind that the designation “future” is now entirely arbitrary and would be
3.5. Spacetime Manifolds and Spinor Structures 171
with group L+↑ . As we argued in Section 2.4 (pages 114–115) the object re-
quired to describe wavefunctions of spin 12 particles on X is a lift of this bundle
to a principal SL(2, C)-bundle over X. More precisely, a spinor structure
for X consists of a principal SL(2, C)-bundle
P
SL(2, C) ,→ S(X) −→
S
X
over X and a map λ : S(X) −→ L(X) such that PL (λ(p)) = PS (p) and
λ(p · g) = λ(p) · Spin(g) for all g ∈ SL(2, C). In Section 6.5 we will determine
a necessary and sufficient condition (on the topology of X) for the existence
of such a spinor structure.
We will conclude this section with a number of examples of spacetime man-
ifolds. We have been guided in the selection of these examples by the desire
to illustrate the concepts we have introduced in a context as free of techni-
cal obfuscations as possible without wandering into the realm of physically
meaningless examples contrived solely for pedegogical purposes. Some of the
examples are of great physical significance, while others are primarily of histor-
ical interest, but all of them have played a role in the development of general
relativity.
172 3. Frame Bundles and Spacetimes
( )
∂ ∂
gαβ (p) = g p , = ηαβ , α, β = 0, 1, 2, 3,
∂xα p ∂xβ p
where
1, α=β=0
ηαβ = −1, α = β = 1, 2, 3 .
0, α ̸= β
Thus, the gαβ are constant and, identifying each Tp (R1,3 ) with R4 via the
canonical isomorphism, we have, in effect, just introduced the Minkowski inner
product on R4 . Relative to the basis {dxα ⊗ dxβ : α, β = 0, 1, 2, 3} for the
covariant tensors of rank 2 on R1,3 , the metric for R1,3 is given by
∑
3
g = ηαβ dxα ⊗ dxβ = dx0 ⊗ dx0 − dxi ⊗ dxi .
i=1
To ease the typography we will often write the coordinate velocity vector
fields ∂x∂α as ∂α , α = 0, 1, 2, 3. The standard orientation for R1,3 is the
one that assigns to each p ∈ R1,3 the orientation for Tp (R1,3 ) contain-
ing {∂0 (p), ∂1 (p), ∂2 (p), ∂3 (p)} and we will time orient R1,3 with the vector
field ∂0 . Thus, a tangent vector v = v α ∂α (p) ∈ Tp (R1,3 ) which is either time-
like (g p (v , v ) = (v 0 )2 − (v 1 )2 − (v 2 )2 − (v 3 )2 > 0) or null (g p (v , v ) = (v 0 )2 −
(v 1 )2 − (v 2 )2 − (v 3 )2 = 0) and nonzero is future directed if g p (∂0 (p), v ) = v 0
is positive. Since any principal bundle over R4 is necessarily trivial, this is
true, in particular, for the oriented, time oriented, orthonormal frame bundle
of R1,3 .
Exercise 3.5.1 Introduce spherical coordinates (t, ρ, ϕ, θ) on R1,3 as follows:
Define a map from
to R1,3 by
x0 = t
x1 = ρ sin ϕ cos θ
x2 = ρ sin ϕ sin θ
x3 = ρ cos ϕ.
3.5. Spacetime Manifolds and Spinor Structures 173
Show that the Jacobian of the map is ρ2 sin ϕ. Use this fact to show that the
map is a diffeomorphism onto an open set in R1,3 . What open set? The inverse
of this map is therefore a chart on R1,3 . Denote the coordinate velocity fields
for this chart ∂t , ∂ρ , ∂ϕ and ∂θ and compute the components of the metric in
this chart (i.e., g (∂t , ∂t ), g (∂t , ∂ρ ), etc.) to show that
Finally, alter the domain of the mapping in such a way as to obtain charts
that cover as much of R1,3 as possible.
Exercise 3.5.2 Introduce advanced and retarded null coordinates v and w
on R1,3 by letting v = t + ρ and w = t − ρ (so that v ≥ w). Thus,
1
t= (v + w)
2
1
ρ = (v − w)
2
ϕ=ϕ
θ = θ.
Show that the Jacobian of the map is identically equal to 12 so that the trans-
formation (v, w, ϕ, θ) −→ (t, ρ, ϕ, θ) is nonsingular wherever it is defined. Show
that, in these coordinates, the metric is given by
1
g = dv ⊗ dw − (v − w)2 (dϕ ⊗ dϕ + sin2 ϕ dθ ⊗ dθ).
4
basis. We will define the Lorentz metric g for E by giving its components gαβ
relative to this global chart. Specifically, at each p = (x0 , x1 , x2 , x3 ) ∈ E,
1 , α=β=0
gαβ (p) = gαβ (x , x , x , x ) = −(x0 )4/3 ,
0 1 2 3
α = β = 1, 2, 3. (3.5.2)
0 , α ̸= β
Thus,
∑
3
g = dx0 ⊗ dx0 − (x0 )4/3 dxi ⊗ dxi (3.5.3)
i=1
b 0 = ∂0 (p)
b i = (x0 )−2/3 ∂i (p), i = 1, 2, 3.
(v 1 )2 + (v 2 )2 + (v 3 )2 = (x0 )−4/3 (v 0 )2 ,
which one can interpret geometrically as saying that the null cones in E
“get steeper” as p “gets higher,” i.e., as time goes on. Since E is diffeomor-
phic to R4 its oriented, time oriented, orthonormal frame bundle is trivial.
Remark: The physical significance of the Einstein-deSitter spacetime as
a cosmological model is discussed at great length in [SaW]. This is not our
concern here so we will content ourselves with a few observations on the proper
way to view E without any real attempt at justification.
The vertical straight lines α(t) = (t, x10 , x20 , x30 ), where 0 < t < ∞
and (x10 , x20 , x30 ) ∈ R3 is fixed, are clearly future directed, timelike curves
in E. These are to be interpreted as the worldlines of the galaxy clus-
ters of our universe. The “displacement vector” v between two events with
the same x0 on two different vertical worldlines is spacelike and satis-
1
fies ∥v ∥ = (−⟨v , v ⟩) 2 = K(x0 )2/3 , where K is a positive constant (here
3.5. Spacetime Manifolds and Spinor Structures 175
(x3 )2 − (x4 )2 is obviously smooth and has −1 as a regular value so Q−1 (−1)
is a 4-dimensional smooth submanifold of R1,4 . As a manifold, D is just this
smooth submanifold of R1,4 , i.e.,
{
D = (x0 , x1 , x2 , x3 , x4 ) ∈ R1,4 :
}
(x1 )2 + (x2 )2 + (x3 )2 + (x4 )2 − (x0 )2 = 1 .
)
(1 + (x0 )2 )− 2 x3 , (1 + (x0 )2 )− 2 x4
1 1
grad Q(p) = 2x0 ∂0 (p) − 2x1 ∂1 (p) − 2x2 ∂2 (p) − 2x3 ∂3 (p) − 2x4 ∂4 (p)
Show that V (p) ∈ Tp (D) and g (V (p), V (p)) > 0. Conclude that D is time
oriented by V = ∂0 + g̃(U , ∂0 )U .
Exercise 3.5.7 Use the oriented, orthonormal frame field {V 1 , V 2 , V 3 } for
S 3 constructed in Section 3.3 (see (3.3.9)) and the timelike, future directed
vector field V on D constructed in Exercise 3.5.6 to show that the oriented,
time oriented, orthonormal frame bundle of D is trivial.
3.5. Spacetime Manifolds and Spinor Structures 177
Then (x1 )2 +(x2 )2 +(x3 )2 +(x4 )2 −(x0 )2 = 1 so the image of τ is in D. For each
fixed t in (−∞, ∞), x1 , x2 , x3 and x4 parametrize the 3-sphere of radius cosh t
and covers the entire 3-sphere for ξ, ϕ and θ restricted to 0 ≤ ξ ≤ π, 0 ≤ ϕ ≤ π
and 0 ≤ θ ≤ 2π. With these values of ξ, ϕ and θ and −∞ < t < ∞, τ maps
onto D.
Exercise 3.5.8 Compute the 5 × 4 Jacobian matrix of τ and show that it
has rank 4 on each of the following regions in R4 .
Show also that τ is one-to-one when restricted to either of these regions and
conclude that the inverse of each of these restrictions is a chart on D.
Exercise 3.5.9 Show that, relative to the coordinates ξ, ϕ, θ, t on D, the
metric g takes the form
( )
g = dt ⊗ dt − cosh2 t dξ ⊗ dξ + sin2 ξ(dϕ ⊗ dϕ + sin2 ϕdθ ⊗ dθ) .
Exercise 3.5.12 Show that C is orientable and time orientable and that its
oriented, time oriented, orthonormal frame bundle is trivial.
Exercise 3.5.13 Show that one can introduce coordinates ξ, ϕ, θ, t on C by
x0 = t
x1 = sin ξ sin ϕ cos θ
x2 = sin ξ sin ϕ sin θ
x3 = sin ξ cos ϕ
x4 = cos ξ
Introduction
Physics is expressed in the language of differential equations (e.g., Maxwell,
Dirac, Yang-Mills, Einstein, etc.). Differential equations live on differentiable
manifolds and differentiable manifolds have topologies that influence not only
the solutions to differential equations defined on them, but even the type of
equation that one can define on them. At some naive level then it is perhaps
not surprising that topology and physics interact. The profound depth of this
interaction in recent years, however, has made it abundantly clear that the
naive level is not the appropriate one from which to view this.
It is remarkable that the deep connection between topology and the differ-
ential equations of physics can be made quite explicit. The bridge between
the two subjects is the notion of an elliptic complex of differential operators
on a manifold and its corresponding cohomology. There is, in fact, an explicit
formula (the Atiyah-Singer Index Theorem) relating the analytic properties of
the differential operators in such a complex to the topology of the underlying
manifold. This is, unfortunately, quite beyond our level here, although the
object will arise again in the Appendix. We will, however, in this chapter con-
struct the simplest example of an elliptic complex (the de Rham complex) and,
in the next, study its cohomology. In Chapter 6 we will use the information
thus accumulated to construct the characteristic classes that have, at least
informally, put in an appearance in our earlier discussions of electromagnetic
and Yang-Mills fields (Chapter 2).
and
A(v1 , . . . , avi , . . . , vk ) = aA(v1 , . . . , vi , . . . , vk )
for all v1 , . . . , vi , vi′ , . . . , vk in E. The set T k (E) of all such multilinear forms
is a real vector space with pointwise operations:
Note that T 1 (E) is the dual space E ∗ . For convenience we will take T 0 (E) =
R. The elements of T k (E) are called covariant tensors of rank k (or simply
k -tensors) on E. If T : E1 −→ E2 is a linear transformation we define the
pullback map
T ∗ : T k (E2 ) −→ T k (E1 )
Exercise 4.1.1 Show that if A ∈ T k (E), then the following three conditions
are equivalent:
(a) A is zero whenever two of its arguments are equal, i.e., if 1 ≤ i,
j ≤ k and i ̸= j, then A(v1 , . . . , vi , . . . , vj , . . . , vk ) = 0 whenever
vi = vj .
(b) A changes sign whenever two of its arguments are interchanged (and
the remaining arguments are left fixed), i.e., A(v1 , . . . ,
vj , . . . , vi , . . . , vk ) = −A(v1 , . . . , vi , . . . , vj , . . . , vk ).
(c) If σ ∈ Sk is any permutation of {1, . . . , k} and (−1)σ is its sign (1 if
σ is an even permutation and −1 if σ is an odd permutation), then
A(vσ(1) , . . . , vσ(k) ) = (−1)σ A(v1 , . . . , vk ).
Exercise 4.1.2 Show that A ⊗ B is, indeed, in T k+l (E) and prove each of
the following properties of the tensor product.
(a) (A1 + A2 ) ⊗ B = A1 ⊗ B + A2 ⊗ B
(b) A ⊗ (B1 + B2 ) = A ⊗ B1 + A ⊗ B2
(c) (aA) ⊗ B = A ⊗ (aB) = a(A ⊗ B) (a ∈ R)
(d) A ⊗ (B ⊗ C) = (A ⊗ B) ⊗ C
(e) T ∗ (A ⊗ B) = (T ∗ A) ⊗ (T ∗ B).
(k + l)!
α∧β = Alt(α ⊗ β). (4.1.3)
k! l!
Lemma 4.1.2 Let k ≥ 0 be an integer. Then
1. A ∈ T k (E) ⇒ Alt(A) ∈ Λk (E).
2. α ∈ Λk (E) ⇒ Alt(α) = α.
3. A ∈ T k (E) ⇒ Alt(Alt(A)) = Alt(A).
Proof: (3) obviously follows from (1) and (2). To prove (1) we observe first
that Alt(A) is clearly multilinear so Alt(A) ∈ T k (E). To prove that Alt(A)
is skew-symmetric we verify (b) in Exercise 4.1.1. Thus, we fix i and j with
1 ≤ i, j ≤ k and show that interchanging the ith and j th arguments changes
the sign of Alt(A). Assume, without loss of generality, that 1 ≤ i < j ≤ k and
let (ij) be the permutation of {1, . . . , k} that switches i and j, but leaves the
others fixed. Thus, for any k -tuple (a, b, . . . , c), (ij) · (a, b, . . . , c) has the ith
and j th slots switched while the others are left fixed. Moreover, as σ varies
′
over all the permutations in Sk , so does σ ′ = (ij) ◦ σ, but (−1)σ = −(−1)σ .
Thus,
4.1. Multilinear Algebra 183
Alt(A)(v1 , . . . , vj , . . . , vi , . . . , vk )
1 ∑ ( )
= (−1)σ A σ · (v1 , . . . , vj , . . . , vi , . . . , vk )
k! σ
( )
1 ∑ ( )
= (−1) A σ · (ij) · (v1 , . . . , vi , . . . , vj , . . . , vk )
σ
k! σ
1 ∑ ( )
= (−1)σ A ((ij) ◦ σ) · (v1 , . . . , vi , . . . , vj , . . . , vk )
k! σ
1 ∑ ′
( )
=− (−1)σ A σ ′ · (v1 , . . . , vi , . . . , vj , . . . , vk )
k! σ
1 ∑ ′
( )
=− (−1)σ A σ ′ · (v1 , . . . , vi , . . . , vj , . . . , vk )
k! ′
σ
= −Alt(A)(v1 , . . . , vi , . . . , vj , . . . , vk ).
1
Remark: Notice that the factor k! played no role in the proof of (1). This
is not the case, however, for the proof (2).
To prove (2) we begin with an α ∈ Λk (E). Observe first that, for any σ ∈ Sk ,
( )
α σ · (v1 , . . . , vk ) = (−1)σ α (v1 , . . . , vk ).
Λl (E), then α ∧ β ∈ Λk+l (E). To establish the basic properties of the wedge
product we need two more technical results.
Lemma 4.1.3 Let A ∈ T k (E) and B ∈ T l (E) and suppose Alt(A) = 0. Then
Alt(A ⊗ B) = Alt(B ⊗ A) = 0.
Proof: Begin by writing
σ0 ◦ G = {σ0 ◦ σ : σ ∈ G}.
Let σ0 · (v1 , . . . , vk+l ) = (w1 , . . . , wk , wk+1 , . . . , wk+l ). Then this last sum
becomes
∑ ( ) ( )
(−1)σ0 (−1)σ A wσ(1) , . . . , wσ(k) B wσ(k+1) , . . . , wσ(k+l)
σ∈G
Since Sk+l is finite we may continue in this way, splitting Sk+l into finitely
many disjoint subsets, the sum over each being zero. Thus,
∑ ( )
(−1)τ (A ⊗ B) τ · (v1 , . . . , vk+l ) = 0
τ ∈ Sk+l
4.1. Multilinear Algebra 185
1. (α1 + α2 ) ∧ β = α1 ∧ β + α2 ∧ β
2. α ∧ (β 1 + β 2 ) = α ∧ β 1 + α ∧ β 2
4. β ∧ α = (−1)kl α ∧ β
(k+l+m)!
5. (α ∧ β) ∧ γ = α ∧ (β ∧ γ) = k! l! m! Alt(α ⊗ β ⊗ γ)
Proof: (1), (2) and (3) follow directly from the corresponding properties of
the tensor product (Exercise 4.1.2). To prove (4) we write
and
(β ∧ α) (v1 , . . . , vk+l )
1 ∑ ( )
= (−1)σ (α ⊗ β) vσ(l+1) , . . . , vσ(l+k) , vσ(1) , . . . , vσ(l)
k! l!
σ ∈ Sk+l
1 ∑ ( ( ))
= (−1)σ (α ⊗ β) ρ · vσ(1) , . . . , vσ(l) , vσ(l+1) , . . . , vσ(l+k)
k! l!
σ ∈ Sk+l
1 ∑ ( ( ))
= (−1)σ (α ⊗ β) ρ · σ · (v1 , . . . , vk+l )
k! l!
σ ∈ Sk+l
1 ∑ ( )
= (−1)kl (−1)σ ◦ ρ (α ⊗ β) (σ ◦ ρ) · (v1 , . . . , vk+l )
k! l!
σ ∈ Sk+l
1 ∑ ′
( )
= (−1)kl (−1)σ (α ⊗ β) σ ′ · (v1 , . . . , vk+l )
k! l!
σ ′ ∈ Sk+l
as required.
β ∧ α = −α ∧ β.
4.1. Multilinear Algebra 187
Then α′ = c det for some constant c. Evaluating α′ at ((1, 0, . . . , 0), . . . , (0, 0, . . . , 1))
gives α′ ((1, 0, . . . , 0), . . . , (0, 0, . . . , 1)) = c det (id), i.e., α(e1 , . . . , en ) = c.
Thus,
α(v1 , . . . , vn ) = α(Ai 1 ei , . . . , Ai n ei )
( )
= α′ (A1 1 , . . . , An 1 ), . . . , (A1 n , . . . , An n )
= α(e1 , . . . , en ) det(Ai j ).
4.1. Multilinear Algebra 189
Corollary 4.1.9 Let {e1 , . . . , en } be a basis for E, {e1 , . . . , en } its dual basis
for Λ1 (E) and v1 , . . . , vn vectors in E with vj = Ai j ei , j = 1, . . . , n. Then
T ∗ α = (det T )α
so
cα (e1 , . . . , en ) = det(Ai j ) α (e1 , . . . , en ).
Since α ̸= 0, Theorem 4.1.8 implies that α(e1 , . . . , en ) ̸= 0 so
c = det(Ai j ).
As another application of Theorem 4.1.8 we show that any nonzero element
of Λn (E) determines a unique orientation for E.
Theorem 4.1.11 Let E be an n-dimensional real vector space and α a
nonzero element of Λn (E). Then there is a unique orientation µ for E such
that [e1 , . . . , en ] ∈ µ if and only if α(e1 , . . . , en ) > 0.
Proof: Since α is nonzero, Theorem 4.1.8 implies that α(e1 , . . . , en ) is
nonzero for every basis {e1 , . . . , en } for E. Thus, the ordered bases for E
are divided into two disjoint classes according to whether α(e1 , . . . , en ) > 0
or α(e1 , . . . , en ) < 0. Let µ denote the set of all ordered bases for E for
which α(e1 , . . . , en ) > 0. We claim that µ is an orientation for E, i.e., that
µ is an equivalence class of the equivalence relation ∼ defined on ordered
bases as follows: {ê1 , . . . , ên } ∼ {e1 , . . . , en } if and only if det(Ai j ) > 0,
where êj = Ai j ei , j = 1, . . . , n. But this is clear from Theorem 4.1.8 since
α(ê1 , . . . , ên ) = det(Ai j )α(e1 , . . . , en ). Uniqueness is also clear since E has
precisely two orientations and the other one (i.e., −µ) consists of the ordered
bases for which α(e1 , . . . , en ) < 0.
Exercise 4.1.13 Show that if {e1 , . . . , en } is an ordered basis for E and
{e1 , . . . , en } is its dual basis for Λ1 (E), then the orientation for E determined
by e1 ∧ · · · ∧ en is precisely the one that contains {e1 , . . . , en }.
190 4. Differential Forms and Integration
Exercise 4.1.14 Show that ω is uniquely determined, i.e., that there is only
one element of Λn (E) that carries every oriented orthonormal basis for E to 1.
Theorem 4.1.12 Let E be an n-dimensional real vector space with an orien-
tation µ and an inner product g. Then there exists a unique
ω ∈ Λn (E) such that ω(e1 , . . . , en ) = 1 whenever {e1 , . . . , en } is an oriented
orthonormal basis for E.
For example, if E = Rn with its standard orientation and (positive definite)
inner product, then ω is just the determinant function det ∈ Λn (Rn ) (see
the first few lines in the proof of Theorem 4.1.8). In general, ω is called
the (metric) volume form for E determined by µ and g (recall that, in
R3 , | det(v1 , v2 , v3 )| is the volume of the parallelepiped spanned by v1 , v2 and
v3 ). We know that if {e1 , . . . , en } is an oriented orthonormal basis for E, then
ω = e1 ∧· · ·∧en , where {e1 , . . . , en } is the dual basis. We will need to compute
ω in an arbitrary oriented (but not necessarily orthonormal) basis for E.
Exercise 4.1.15 Let ω be the volume form on E determined by the orienta-
tion µ and an inner product g. Let {ê1 , . . . , ên } be an oriented basis for E and
{ê1 , . . . , ên } its dual basis. For each i, j = 1, . . . , n let ĝij = g(êi , êj ). Show that
1
ω = | det(ĝij )| 2 ê1 ∧ · · · ∧ ên .
the matrix inverse of (Ai j ). Then the dual bases {e1 , . . . , en } and {ê1 , . . . , ên }
are related by êj = Ai j ei , j = 1, . . . , n.
Exercise 4.1.16 Let ĝij = g(êi , êj ) and let (ĝ ij ) be the matrix inverse of
(ĝij ). Show that
ĝij = Ak i Al j gkl
and
ĝ ij = Ak i Al j g kl
for i, j = 1, . . . , n.
1 1
Now, for α ∈ Λk (E) we write α = αi ···i ei1 ∧ · · · ∧ eik = α̂i ...i êi1
k! 1 k k! 1 k
∧ · · · ∧ êik , where
( )
α̂i1 ...ik = α(êi1 , . . . , êik ) = α Aj1 i1 ej1 , . . . , Ajk ik ejk
= Aj1 i1 · · · Ajk ik αj1 ···jk .
Exercise 4.1.17 Show that α̂j1 ...jk = Al1 j1 . . . Alk jk αl1 ···lk .
Finally, we compute
1 j1 ···jk 1 ( j1 )( )
α̂ β̂j1 ···jk = Al1 · · · Alk jk αl1 ···lk Am1 j1 · · · Amk jk βm1 ···mk
k! k!
1 ( j 1 m1 ) ( )
= Al1 A j1 · · · Alk jk Amk jk αl1 ···lk βm1 ···mk
k!
1
= δl1m1 · · · δlkmk αl1 ···lk βm1 ···mk
k!
1
= αl1 ···lk βl1 ···lk
k!
1
= αj1 ···jk βj1 ···jk
k!
as required. Thus, g is well-defined on Λk (E) and it is clearly bilinear, sym-
metric and nondegenerate.
Theorem 4.1.13 Let E be an n-dimensional real vector space with an in-
ner product g and let k be an integer with 1 ≤ k ≤ n. If {e1 , . . . , en }
is an orthonormal basis for E with dual basis {e1 , . . . , en } for Λ1 (E), then
{ei1 ∧ · · · ∧ eik : 1 ≤ i1 < · · · < ik ≤ n} is an orthonomal basis for Λk (E)
relative to the induced metric g on Λk (E) (defined by (4.1.5)).
1 m1 ···mk 1
g(α, β) = α βm1 ···mk = g l1 m1 · · · g lk mk αl1 ···lk βm1 ···mk .
k! k!
Notice that this will be zero unless l1 = m1 , . . . , lk = mk so
1 l1 l1
g(α, β) = g . . . g lk lk αl1 ···lk βl1 ···lk .
k!
In order for αl1 ···lk to be nonzero, {l1 , . . . , lk } must equal {i1 , . . . , ik }. Simi-
larly, βl1 ···lk will be zero unless {l1 , . . . , lk } = {j1 , . . . , jk }. Thus, g(α, β) will
be zero unless {i1 , . . . , ik } = {j1 , . . . , jk }. In particular,
Remark: The last few lines of this proof actually show that
g(ei1 ∧ · · · ∧ eik , ei1 ∧ · · · ∧ eik ) = (−1)m ,
where m is the number of indices among {i1 , . . . , ik } for which gii = −1.
Theorem 4.1.14 Let E be an n-dimensional real vector space with an ori-
entation µ and an inner product g. Let ω be the metric volume form for E
detemined by µ and g and let k be an integer with 0 ≤ k ≤ n. Then there
exists a unique isomorphism
∗
: Λk (E) −→ Λn−k (E)
such that
α ∧ ∗ β = g(α, β)ω (4.1.6)
for all α, β ∈ Λ (E).
k
φγ : Λk (E) −→ R
194 4. Differential Forms and Integration
Some caution is in order here, however, since we have not assumed that our
inner product is positive definite so that ∥β∥2 need not be positive.
4.1. Multilinear Algebra 195
We will need some formulas for computing Hodge duals. These are
particularly simple relative to an oriented orthonormal basis {e1 , . . . , en } so we
will consider these first. First observe that if β ∈ Λ0 (E) = R, then ∗ β ∈ Λn (E)
and so is a real multiple of ω = e1 ∧ · · · ∧ en .
Exercise 4.1.19 Show that, if β ∈ Λ0 (E) = R, then
∗
β == βω = βe1 ∧ · · · ∧ en ,
∗
(e1 ∧ · · · ∧ en ) = (−1)s 1,
∗
(ei1 ∧ · · · ∧ ein−1 ) = ±ein
where {i1 , . . . , in−1 , in } = {1, . . . , n} and one chooses the plus (minus) sign if
i1 · · · in−1 in is an even (odd) permutation of 1· · · n.
Finally, if k < n − 1 one can select l1 , . . . , ln−k so that i1 · · · ik l1 · · · ln−k is an
even permutation of 1 · · · n. Then
so
∗
(ei1 ∧ · · · ∧ eik ) = (−1)m el1 ∧ · · · ∧ eln−k , (4.1.8)
where, again, m is the number of indices i1 , . . . , ik with gii = −1.
Exercise 4.1.21 Verify the following concrete examples:
(a) Let E = R3 with its standard orientation and (positive definite)
inner product and let {e1 , e2 , e3 } be an oriented orthonormal basis.
196 4. Differential Forms and Integration
Then
∗ 1 ∗ 2 ∗ 3
e = e2 ∧ e3 e = −e1 ∧ e3 e = e1 ∧ e2
∗ ∗ ∗
(e1 ∧ e2 ) = e3 (e2 ∧ e3 ) = e1 (e1 ∧ e3 ) = −e2
∗
(e1 ∧ e2 ∧ e3 ) = 1.
(c) Let E = R1,3 with its standard orientation and Minkowski inner product
and let {e0 , e1 , e2 , e3 } be an oriented orthonormal basis
(assume, as usual, that g(e0 , e0 ) = 1 and g(ei , ei ) = −1 for i = 1, 2, 3).
Then
∗ 2 ∗
e = −e0 ∧ e1 ∧ e3 (e1 ∧ e2 ) = e0 ∧ e3
∗ ∗
(e ∧ e ∧ e ) = −e
0 1 3 2
(e0 ∧ e1 ∧ e2 ∧ e3 ) = −1.
To describe the Hodge dual in an arbitrary oriented basis {ê1 , . . . , ên } with
dual basis {ê1 , . . . , ên } we will use the Levi-Civita symbol
1, if j1 · · · jn is an even permutation of 1 · · · n
εj1 ···jn = −1, if j1 · · · jn is an odd permutation of 1 · · · n
0, otherwise
and borrow a result from the theory of determinants: For any n × n matrix
A = (Ai j )i,j=1,...,n ,
Now let {e1 , . . . , en } be an oriented orthonormal basis for E with dual basis
{e1 , . . . , en }. We write êj = Ai j ei , j = 1, . . . , n, and denote by (Ai j ) the in-
verse of the matrix (Ai j ). Then êj = Ai j ei . In addition, we write gij = g(ei , ej )
and ĝij = g(êi , êj ) for i, j = 1, . . . , n, denote by (g ij ) and (ĝ ij ) the inverses
of the matrices (gij ) and (ĝij ), respectively, and recall (from Exercise 4.1.15)
that
1
det(Ai j ) = | det(ĝij )| 2 . (4.1.10)
4.1. Multilinear Algebra 197
the order i1 , . . . , ik ). There are s − m indices among {l1 , . . . , ln−k } for which
gll = −1. Then
(
∗ l1
)
e ∧ · · · ∧ eln−k = (−1)s−m (−1)k(n−k) ei1 ∧ · · · ∧ eik
so m′ = k(n − k) + s − m. Thus
∗∗
β = (−1)m (−1)s−m (−1)k(n−k) ei1 ∧ · · · ∧ eik = (−1)k(n−k)+s β
as required.
Corollary 4.1.16 If E is an oriented vector space with a positive definite in-
ner product g, then each Hodge star isomorphism ∗: Λk (E) −→ Λn−k (E), 0 ≤
k ≤ n, is an isometry, i.e.,
we discuss in some detail, virtually all of the material in Section 4.1 general-
izes immediately to this context by simply doing everything componentwise
with respect to some basis for V.
As before we let E denote some n-dimensional real vector space. For some
purposes (to be specified as we proceed), E will be assumed to have an orien-
tation µ and an inner product g. Our forms will be defined on E and will take
values in some m-dimensional real vector space V. When appropriate we will
also assume that V has an inner product h (e.g., when V is a Lie algebra, h
will generally arise from the Killing form). We will use {e1 , . . . , en } to denote
a generic basis for E and {T1 , . . . , Tm } will be a basis for V. A map
k
A : E × · · · × E −→ V
T ∗ : T k (E2 , V) −→ T k (E1 , V)
by
( )
(T ∗ A)(v1 , . . . , vk ) = A T (v1 ), . . . , T (vk )
for any A ∈ T k (E2 , V).
If {T1 , . . . , Tm } is any basis for V, then any A ∈ T k (E, V) can be written
uniquely as
A = A1 T1 + · · · + Am Tm = Ai Ti ,
componentwise:
∗ ∗
α= (α1 T1 + · · · + αm Tm ) = ∗α1 T1 + · · · + ∗αm Tm . (4.2.1)
Exercise 4.2.1 Show that this definition does not depend on the choice of
basis for V.
If we assume that V also has an inner product h, then, together with g and
the induced inner products on each Λk (E) (also denoted g) we can define
inner products, denoted (gh), on each Λk (E, V) as follows: Let {T1 , . . . , Tm }
be a basis for V, write α, β ∈ Λk (E, V) as α = αi Ti and β = β j Tj and set
hij = h(Ti , Tj ) for i, j = 1, . . . , m. Now define
Exercise 4.2.2 Show that this definition does not depend on the choice of
basis for V.
Exercise 4.2.3 Show that, if α, β ∈ Λ0 (E, V) = V, then (gh) coincides with h.
( )( )
A ⊗ρ B v1 , . . ., vk , vk+1 , . . ., vk+l
( ) (4.2.3)
= ρ A (v1 , . . . , vk ), B(vk+1 , . . . , vk+l ) .
Exercise 4.2.5 Verify that A ⊗ρ B is, indeed, in T k+l (E, W) and that ⊗ρ
has all of the properties of ⊗ described in Exercise 4.1.2.
If α ∈ Λk (E, U) and β ∈ Λl (E, V), then their ρ-wedge product α ∧ρ β ∈
Λk+l (E, W) is defined by
( )( )
α ∧ρ β v1 , . . . , vk+l
1 ∑ ( )( ) (4.2.4)
= (−1)σ α ⊗ρ β vσ(1) , . . . , vσ(k+l) ,
k!l! σ
α ∧ρ β = (α1 + α2 i ) ∧ρ (β 1 + β 2 i )
= (α1 ∧ β 1 − α2 ∧ β 2 ) + (α1 ∧ β 2 + α2 ∧ β 1 )i .
ρ(A, B) = [A, B] = AB − BA
α ∧ρ β = [α, β]
for any G-valued forms α and β and we will adhere to this custom.
Thus, for any α ∈ Λk (E, G) and β ∈ Λl (E, G) and any v1 , . . . , vk+l in E
we have
( )
[α, β] v1 , . . . , vk+l
(4.2.5)
1 ∑ [ ( ) ( )]
= (−1)σ α vσ(1) , . . . , vσ(k) , β vσ(k+1) , . . . , vσ(k+l) .
k!l!
σ∈Sk+l
[α, β] = [αi Ti , β j Tj ]
(4.2.6)
= (αi ∧ β j )[Ti , Tj ] = Cij
k
(αi ∧ β j )Tk .
that
( )
(α ∧ β) v1 , . . . , vk+l
1 ∑ ( ) ( ) (4.2.7)
σ
= (−1) α vσ(1) , . . . , vσ(k) β vσ(k+1) , . . . , vσ(k+l) .
k! l!
σ ∈Sk+l
as promised.
We mention one concrete example that has already come up in the context
of SU (2)-Yang-Mills-Higgs theory in Section 2.5. Let E be any n-dimensional
real vector space with an orientation µ and an innerproduct g. For V we take
the Lie algebra su(2) of all 2 × 2 skew-Hermitian, tracefree matrices. We
have defined a (positive definite) inner product h on su(2) by h(A, B) = −2
trace(AB). The basis {T1 , T2 , T3 } for su(2) consisting of
1 1 1
T1 = − i σ1 T2 = − i σ2 T3 = − i σ3 ,
2 2 2
where
( ) ( ) ( )
0 1 0 −i 1 0
σ1 = σ2 = σ3 =
1 0 i 0 0 −1
are the Pauli spin matrices is h-orthonormal, i.e., hij = h(Ti , Tj ) = δij for
i, j = 1, 2, 3. If α, β ∈ Λk (E, su(2)), then
3 2 1
1 α i α + α i
α = α1 T1 + α2 T2 + α3 T3 = − 2 (4.2.10)
2 −α + α1 i −α3 i
so, in particular,
1 ( ) 1
−trace αi1 ···ik αi1 ···ik = ∥α∥2 .
k! 2
Combining this with Exercise 4.2.10 gives
1 1 ( )
−trace(α ∧ ∗ α) = ∥α∥2 ω = − trace αi1 ···ik αi1 ···ik ω. (4.2.11)
2 k!
component functions αi1 ···ik (x) are C ∞ on U for all charts in some atlas for
X. The set of all such is denoted Λk (X) (with Λ0 (X) = C ∞ (X)) and has the
obvious pointwise structure of a C ∞ (X)-module.
k
2. Every α ∈ Λk (X) gives rise to a map α : X (X) × · · · × X (X) −→
C ∞ (X) defined by α(V1 , . . . ,Vk )(x) = α(x)(V 1 (x), . . . , V k (x)) that is skew-
symmetric and C ∞ (X)-multilinear. Conversely, any skew-symmetric, C ∞ (X)-
k
multilinear map A : X (X) × · · · × X (X) −→ C ∞ (X) determines a unique
α ∈ Λk (X) with α(V1 , . . . ,Vk ) = A(V1 , . . . ,Vk ) and this one-to-one corre-
spondence is a C ∞ (X)-module isomorphism. Thus, a differential k -form
α on X can be identified with a skew-symmetric, C ∞ (X)-multilinear map
α : X (X) × · · · × X (X) −→ C ∞ (X).
3. Let GL(n, R) ,→ L(X) −→ X be the linear frame bundle of X. Let
V be the real vector space Λk (Rn ) of skew-symmetric, k-multilinear forms
on Rn and define a representation ρ : GL(n, R) −→ GL(Λk (Rn )) by
(ρ(g)(α))(v1 , . . . , vk ) = α(g T v1 , . . . , g T vk ), where each vi ∈ Rn is written as a
column matrix and g T vi : is the matrix product. The corresponding associated
vector bundle L(X) ×ρ Λk (Rn ) is called the exterior k -bundle of X and a
differential k -form α on X can be identified with a smooth cross-section
α : X −→ L(X) ×ρ Λk (Rn ), Pρ ◦ α = idX , of L(X) ×ρ Λk (Rn ). Equiva-
lently (Section 6.8, [N4]) a differential k -form α on X can be identified
with a smooth map α : L(X) −→ Λk (Rn ) that is ρ-equivariant, i.e., satisfies
α(p · g) = (ρ(g −1 ))(α(p)) for each g ∈ GL(n, R) and each p ∈ L(X).
Exercise 4.3.1 Show that these three definitions are equivalent in the follow-
ing sense. Let Λk1 (X), Λk2 (X) and Λk3 (X) denote the C ∞ -modules specified by
4.3. Differential Forms 207
Hint: The equivalence of the first two definitions is established for 1-forms on
pages 265–266 of [N4]. The argument hinges on Lemma 5.7.1 of [N4] which
will be needed here as well. For the equivalence of the first and third definitions
proceed as we did for 1-forms in Section 3.3.
Each of these three views of a k-form on a manifold has its uses (unless
some particular emphasis is required we will generally omit the adjective “dif-
ferential”). For example, the purely algebraic operations on forms such as the
wedge product extend immediately to the manifold setting via the first def-
inition by simply doing everything pointwise: If α ∈ Λk (X) and β ∈ Λl (X),
then α ∧ β is the element of Λk+l (X) defined, at each x ∈ X, by
( )
(α ∧ β)(x) (v 1 , . . . , v k+l ) = (α(x) ∧ β(x)) (v 1 , . . . , v k+l )
Exercise 4.3.2 Show that the wedge product is C ∞ (X)-bilinear, i.e., that if
α ∈ Λk (X), and β ∈ Λl (X) and f ∈ C ∞ (X), then
(f α) ∧ β = α ∧ (f β) = f (α ∧ β) .
k dy 1 ∧ · · · ∧ dy n = h dx1 ∧ · · · ∧ dxn
on U ∩ V , then ( )
∂y i
h = k det .
∂xj
Exercise 4.3.4 and Theorem 4.1.11 provide a link between differential forms
and the orientability of a smooth manifold.
Theorem 4.3.1 An n-dimensional manifold X is orientable if and only if X
admits a nowhere zero smooth n-form ω.
Now, the metric for S n is just the restriction to S n of the standard metric
on Rn+1 . The standard orientation for S n (Section 5.10, [N4]) is the one
for which the stereographic projection map φS : US −→ Rn (Example #4,
page 4) is an orientation preserving diffeomorphism (and φN : UN −→ Rn is
orientation reversing). A basis for the tangent space at some point p of S n is
in this orientation if and only if one obtains an oriented basis for Rn+1 by
adjoining to it (at the beginning) an “outward pointing” normal vector to S n
at p.
Exercise 4.3.7 Let p be a point in S n and {e1 , . . . , en } a basis for Tp (S n ) ⊆
Tp (Rn+1 ) = Rn+1 . Show that {e1 , . . . , en } is an oriented orthonormal basis
for Tp (S n ) if and only if {p, e1 , . . . , en } is an oriented orthonormal basis for
Tp (Rn+1 ) = Rn+1 .
210 4. Differential Forms and Integration
Exercise 4.3.7 makes it clear how to define the volume form ω on S n . At each
p ∈ S n and for any v1 , . . . , vn ∈ Tp (S n ) we set
( ) ( )( )
ω p v1 , . . . , vn = dx1 ∧ · · · ∧ dxn+1 p, v1 , . . . , vn
p1 . . . pn+1
1
v1 . . . v1n+1
= det . .. . (4.3.1)
..
.
vn1 . . . vnn+1
∑
ci ∧ · · · ∧ dxn+1 (v1 , . . . , vn )
n+1
= (−1)i−1 pi dx1 ∧ · · · ∧ dx
i=1
(n+1 )
∑ ( )
= ci ∧ · · · ∧ dxn+1
(−1)i−1 xi dx1 ∧ · · · ∧ dx v1 , . . . , vn .
i=1 p
∑
n+1
ω̃ = ci ∧ · · · ∧ dxn+1
(−1)i−1 xi dx1 ∧ · · · ∧ dx (4.3.2)
i=1
4.3. Differential Forms 211
at each point. Thus, one can define the Hodge dual of a k-form on X pointwise:
(∗β)x (v 1 , . . . , v n−k ) = ∗
(β x )(v 1 , . . . , v n−k ).
Exercise 4.3.8 Show that, thus defined, the Hodge dual of a smooth k-form
is a smooth (n − k)-form and that the resulting operator
∗
: Λk (X) −→ Λn−k (X)
is C ∞ (X)-linear.
α ∧ ∗ β = g (α, β)ω,
where g (α, β) is the inner product of the forms α and β defined pointwise by
(4.1.4) and (4.1.5) (and therefore a C ∞ function on X).
We conclude this section by observing that, if V is a finite dimensional
real vector space, then a V-valued differential form α on a manifold X is
defined pointwise in the obvious way and, for any choice of a basis {T1 , . . . , Tm }
for V, can be written α = α1 T1 + · · · + αm Tm , where each αi is a real-valued
differential form on X. All of the algebraic material in Section 4.2 extends at
once to this context by simply doing everything pointwise (the reader who is
skeptical and/or scrupulously honest is encouraged to check all of this out).
212 4. Differential Forms and Integration
∂Θj i
dΘ = d(Θj dxj ) = dΘj ∧ dxj = dx ∧ dxj .
∂xi
Before proceeding with the general definition we ask the reader to take this one
step further.
Exercise 4.4.1 Let Ω ∈ Λ2 (X) be a 2-form on X (thought of as a bilinear
operator on vector fields) and define dΩ : X (X) × X (X) × X (X) −→ C ∞ (X)
by
dΩ(V1 , V2 , V3 ) = V1 (Ω(V2 ,V3 )) − V2 (Ω(V1 ,V3 ))
+ V3 (Ω(V1 ,V2 )) − Ω([V1 ,V2 ],V3 )
+ Ω([V1 ,V3 ],V2 ) − Ω([V2 ,V3 ],V1 ).
d(a1 ω 1 + a2 ω 2 ) = a1 dω 1 + a2 dω 2 . (4.4.1)
4.4. The de Rham Complex 213
d(f ∧ ω) = df ∧ ω + f ∧ dω .
From this and the anti-commutativity of wedge products for 1-forms we obtain
d(ω ∧ f ) = dω ∧ f − ω ∧ df .
Exercise 4.4.2 Let ω 1 ∈ Λk (X) and ω 2 ∈ Λl (X) with k + l ≤ 2. Show that
d(dω) = 0 . (4.4.3)
We show now that there is exactly one way to generalize all of this to higher
degree forms and retain properties (4.4.1), (4.4.2) and (4.4.3).
Theorem 4.4.1 Let X be a smooth n-dimensional manifold. Then there exists
a unique family of operators
( )
1
dω = d ωi1 ···ik dxi1 ∧ · · · ∧ dxik
k!
(4.4.4)
1 ( )
= dωi1 ···ik ∧ dxi1 ∧ · · · ∧ dxik .
k!
214 4. Differential Forms and Integration
1 [( )
= dωi1 ···ik ∧ dxi1 ∧ · · · ∧ dxik
k! ]
+ (−1)0 ωi1 ···ik ∧ d(dxi1 ∧ · · · ∧ dxik ) (by (1) and (2))
1( )
= dωi1 ···ik ∧ dxi1 ∧ · · · ∧ dxik
k!
since each d(dxi1 ∧ · · · ∧ dxik ) = 0 by (2), (3) and induction.
Next we write out an explicit, coordinate independent formula for an op-
erator on Λk (X) which yields elements of Λk+1 (X) and satisfies (1)–(4) (and
so must, therefore, be given in local coordinates by (4.4.4)). Specifically, we
claim that, for each ω ∈ Λk (X), dω must be given by
∑
k+1 ( ( ))
dω(V1 , . . . ,Vk+1 ) = b i , . . . ,Vk+1
(−1)i+1 Vi ω V1 , . . . ,V
i=1
∑ ( ) (4.4.5)
+ b i , . . . ,V
(−1)i+j ω [Vi , Vj ], V1 , . . . ,V b j , . . . ,Vk+1 .
1≤i<j≤n
Exercise 4.4.4 Show that dω ∈ Λk+1 (X). Hint: See Exercise 4.4.1 and its
Hint.
and
( )
∑n
∂f
d(dω) = d dx ∧ dx ∧ · · · ∧ dx
i i1 ik
i=1
∂xi
∑n ( )
∂f
= d dx ∧ dx ∧ · · · ∧ dx
i i1 ik
i=1
∂xi
∑n ( )
∂f
= d ∧ dxi ∧ dxi1 ∧ · · · ∧ dxik
i=1
∂xi
∑n ∑n 2
∂ f
= dxj ∧ dxi ∧ dxi1 ∧ · · · ∧ dxik
∂x j ∂xi
i=1 j=1
∑
n ∑
n
∂2f
= dxj ∧ dxi ∧ dxi1 ∧ · · · ∧ dxik .
i=1 j=1
∂xj ∂xi
Now, every term in this sum with i = j is zero since dxi ∧ dxi = 0 and, when
i ̸= j, the terms
∂2f
dxj ∧ dxi ∧ dxi1 ∧ · · · ∧ dxik
∂xj ∂xi
and
∂2f
dxi ∧ dxj ∧ dxi1 ∧ · · · ∧ dxik
∂xi ∂xj
Exercise 4.4.7 Show that d commutes with pullback. More precisely, let
F : X −→ Y be a smooth map and ω ∈ Λk (Y ). Show that
Hint: This is already known when k = 0 (page 10) and k = 1 (page 12).
Assume the result for (k−1)-forms and prove (4.4.6) for ω = f dxi1 ∧· · ·∧dxik .
Exercise 4.4.8 Let X = R3 with its standard Riemannian metric and ori-
entation and let x, y and z be the standard coordinate functions on R3 (so
that { ∂x
∂ ∂
, ∂y ∂
, ∂z } is a global oriented orthonormal frame field and {dx, dy, dz}
is the dual oriented orthonormal coframe field). Let ∗ denote the correspond-
ing Hodge star operator. Finally, consider the natural isomorphism between
Λ1 (R3 ) and X (R3 ) under which a 1-form α = f dx + g dy + h dz corresponds
∂ ∂ ∂
to the vector field V = f ∂x + g ∂y + h ∂z .
(a) Show that, for any f ∈ Λ0 (R3 ), df ∈ Λ1 (R3 ) corresponds to grad f ∈
X (R3 ).
(b) Show that if α ∈ Λ1 (R3 ) corresponds to V ∈ X (R3 ), then ∗dα ∈
Λ1 (R3 ) corresponds to curl V ∈ X (R3 ).
(c) Show that if α ∈ Λ1 (R3 ) corresponds to V ∈ X (R3 ), then ∗d∗α ∈
Λ0 (R3 ) is divV .
(d) Show that, for any f ∈ Λ0 (R3 ), ∗d ∗df is the Laplacian of f (i.e.,
∗ ∗
d df = ∇2 f = div grad f ).
(e) Show that if α, β ∈ Λ1 (R3 ) correspond to V, W ∈ X (R3 ), then
∗
(α ∧ β) ∈ Λ1 (R3 ) corresponds to V × W ∈ X (R3 ).
(f) Use (3) of Theorem 4.4.1 to prove that curl(grad f ) = 0 and
div(curlV ) = 0.
(g) Use (2) of Theorem 4.4.1 to prove that curl(f V ) = grad f × V +
f curl V .
Remark: The upshot of Exercise 4.4.8 is that the calculus of forms on R3 is
just a disguised version of classical vector analysis. One reason for preferring
forms (aside from the elegant integration theory we construct in Section 4.6)
is that they make sense on higher dimensional manifolds whereas much of
vector calculus (e.g., the cross product and curl) do not. Thus, for example,
once they are written in terms of forms, Maxwell’s equations are defined on
any spacetime.
The exterior differentiation operators are linear transformations on vector
spaces of smooth forms and when all of these are collected into the sequence
d0 d1 dn−2 dn−1 dn
Λ0 (X) −→ Λ1 (X) −→ · · · −→ Λn−1 (X) −→ Λn (X) −→ 0
one obtains what is called the de Rham complex of the n-dimensional mani-
fold X. Theorem 4.4.1 (3) asserts that the composition of any two consecutive
4.4. The de Rham Complex 217
maps in this sequence is identically zero, i.e., the image of any dk−1 is con-
tained in the kernel of dk . A differential form ω on X is said to be closed if
dω = 0 and exact if ω = dη for some form η of degree one less. Thus, one may
rephrase Theorem 4.4.1 (3) by saying that any exact form is closed. The con-
verse is generally not true and we will construct an explicit example shortly.
Indeed, we will devote what remains of this section and all of Chapter 5 to
the issue of when closed implies exact and, when it does not, the extent to
which it does not. Although not apparent at the moment, these are questions
about the topology of X.
We begin our discussion in R2 . If ω is a 1-form on R2 and standard coor-
dinates are x1 and
( ∂ω
2 1 2
) x 1, then2 we can write ω = ω1 dx + ω2 dx and compute
dω = ∂x1 − ∂x2 dx ∧ dx . Thus, if ω is closed we must have
2 ∂ω1
∂ω2 ∂ω1
= .
∂x1 ∂x2
Now define η ∈ Λ0 (R2 ) by
∫ 1 ∫ 1
1 2 1 1 2 2
η(x , x ) = x ω1 (tx , tx )dt + x ω2 (tx1 , tx2 )dt .
0 0
Observe that
∫ 1 ∫ 1
∂η ∂ω1
= x1 (tx 1
, tx 2
)t dt + ω1 (tx1 , tx2 )dt
∂x1 ∂x 1
0
∫ 1 0
2 ∂ω2 1 2
+x 1
(tx , tx )t dt
0 ∂x
∫ 1[ ( )
1 2 1 ∂ω1 1 2
= ω1 (tx , tx ) + t x (tx , tx )
0 ∂x1
( )]
2 ∂ω1 1 2
+t x (tx , tx ) dt
∂x2
∫ 1
d [ ] 1
= t ω1 (tx1 , tx2 ) dt = tω1 (tx1 , tx2 )
0 dt 0
= ω1 (x1 , x2 ).
Similarly,
∂η
= ω2 (x1 , x2 )
∂x2
so
ω = dη .
Thus, on R2 , every closed 1-form is, indeed, exact. We will generalize this
simple result quite substantially when we prove the Poincaré Lemma, but for
the moment we simply wish to contrast it with the situation on R2 − {(0, 0)}.
Here we will write down an explicit 1-form that is closed, but not exact.
218 4. Differential Forms and Integration
We construct our 1-form from two polar angular coordinate functions θ1 and
θ2 . First let L1 = { p ∈ R2 : x1 (p)≥0, x2 (p) = 0 } and U1= R2 − L1 . Define
√1 : U1 −→ R by φ1 (x1 , x2 ) = (r(x , x ), θ1 (x , x )), where r(x , x ) =
2 1 2 1 2 1 2 1 2
φ
1 2 2 2
(x ) + (x ) and θ1 (x , x )) is the unique angle in (0, 2π) such that
tan θ1 (x1 , x2 )) = x2 /x1 . More precisely,
arctan(x2 /x1 ) , x1 > 0, x2 > 0
π
, x1 = 0, x2 > 0
2
θ1 (x1 , x2 ) = π + arctan(x2 /x1 ) , x1 < 0
3π
, x1 = 0, x2 < 0
2
2 π + arctan(x2 /x1 ), x1 > 0, x2 < 0
−x2 1 x1
dθ2 = dx + dx2 on U2 .
(x1 )2 + (x2 )2 (x1 )2 + (x2 )2
Suppose to the contrary that there exists an f ∈ C ∞ (R2 − {(0, 0)}) such
that df = ω on R2 − {(0, 0)}. Then, in particular, on R2 − L1 , df = dθ1 so
d(f − θ1 ) = 0. Thus ∂x 1 = ∂x1 and ∂x2 = ∂x2 on R − L1 . Since R − L1
∂f ∂θ1 ∂f ∂θ1 2 2
hk : Λk (U ) −→ Λk−1 (U ), k ≥ 1,
such that
dk−1 ◦ hk + hk+1 ◦ dk = idΛk (U ) (4.4.7)
ω = f dxi1 ∧ · · · ∧ dxik
(no assumption about the ordering of the indices) and extend by linearity.
Since U is star-shaped with respect to the origin we may define hk ω for such
220 4. Differential Forms and Integration
an ω by
∑
k (∫ 1 )
k
(h ω)(x) = (−1) a−1
t k−1
f (tx)dt xia dxi1 ∧ · · ·
a=1 0
d (4.4.8)
∧ dx ia ∧ · · · ∧ dxik .
∑
k {(∫ 1 )
d k−1 k
(h ω) = (−1) a−1
d k−1
t f (tx)dt xia dxi1 ∧ · · ·
a=1 0
}
d
∧ dx ia ∧ · · · ∧ dxik
∑
k ((∫ 1 ) )
= (−1) a−1
d tk−1
f (tx)dt x ia
∧ dxi1 ∧ · · ·
a=1 0
d
∧ dx ia ∧ · · · ∧ dxik
Thus,
∑
k (∫ 1 )
d k−1 k
(h ω) = (−1) a−1
t k−1
f (tx)dt dxia ∧ dxi1 ∧ · · ·
a=1 0
d
∧ dx ia ∧ · · · ∧ dxik
∑k ∑ n (∫ 1 )
k ∂f
+ (−1) a−1
t j
(tx)dt xia dxj ∧ dxi1 ∧ · · ·
a=1 j=1 0 ∂x
d
∧ dx ia ∧ · · · ∧ dxik
(∫ 1 )
=k tk−1
f (tx)dt dxi1 ∧ · · · ∧ dxik
0
∑
k ∑
n (∫ 1 )
∂f
+ (−1)a−1
t k
(tx)dt xia dxj ∧ dxi1 ∧ · · ·
a=1 j=1 0 ∂xj
d
∧ dx ia ∧ · · · ∧ dxik .
4.4. The de Rham Complex 221
Now,
∑n
∂f
dk ω = j
dxj ∧ dxi1 ∧ · · · ∧ dxik
j=1
∂x
∑
n ∑
k (∫ 1 )
∂f
+ (−1) a
t k
(tx)dt xia dxj ∧ dxi1 ∧ · · ·
j=1 a=1 0 ∂xj
d
∧ dx ia ∧ · · · ∧ dxik .
Thus, computing dk−1 (hk ω) + hk+1 (dk ω), the two double sums cancel and we
obtain
(∫ 1 )
dk−1 (hk ω) + hk+1 (dk ω) = k tk−1 f (tx)dt dxi1 ∧ · · · ∧ dxik
0
n (∫
∑ 1 )
∂f
+ k
t (tx)dt xj dxi1 ∧ · · · ∧ dxik
j=1 0 ∂xj
{∫ 1 [
= ktk−1 f (tx)
0
∑
n
∂f ] }
+ t j
(tx)x dt dxi1 ∧ · · · ∧ dxik
k
j
j=1
∂x
{∫ 1 }
d k
= [t f (tx)]dt dxi1 ∧ · · · ∧ dxik
0 dt
= f (x)dxi1 ∧ · · · ∧ dxik = ω
dω = dω 1 T1 + · · · + dω m Tm
and show that this definition does not depend on the choice of {T1 , . . . , Tm }.
222 4. Differential Forms and Integration
for each g ∈ G.
Proof: We ask the reader to prove uniqueness and the necessity of the two
conditions.
Exercise 4.5.1 Show that projections, when they exist, must be unique (i.e.,
that P ∗ φ̄1 = P ∗ φ̄2 implies φ̄1 = φ̄2 ) and that any φ which does project
to X must satisfy #1 and #2. Note: When k = 0, #1 is taken to be sat-
isfied vacuously and #2 simply says that φ(p · g) = φ(p). In this case the
unique projection φ̄ is obviously given by φ̄(x) = φ(p) for any p ∈ P −1 (x).
Henceforth, we assume k ≥ 1.
Now, suppose φ is a k-form on P satisfying #1 and #2. We define φ̄ on X
as follows: Let x ∈ X and w 1 , . . . , w k ∈ Tx (X). Select some p ∈ P −1 (x) and
then select v 1 , . . . , v k ∈ Tp (P ) with P∗p (v i ) = w i for each i = 1, . . . , k. Now
define
φ̄x (w 1 , . . . , w k ) = φp (v 1 , . . . , v k ). (4.5.1)
224 4. Differential Forms and Integration
Remark: We will prove next that the definition (4.5.1) does not depend
on the choices of p and v 1 , . . . , v k . First, however, we point out the most
convenient way to actually make these choices in practice. Choose a local
cross-section s : V −→ P −1 (V ) with x ∈ V . Let p = s(x) and v i = s∗x (w i )
for i = 1, . . . , k. Then P(p) = P(s(x)) = x and P∗p (v i ) = P∗s(x) (s∗x (w i )) =
(P ◦ s)∗x (w i ) = ( idV )∗x (w i ) = w i as required. Notice that (4.5.1) now gives
so
φ̄ = s∗ φ (4.5.2)
on V .
To see that our definition is independent of the choices suppose that p′ ∈
P −1 (x) and v ′1 , . . . , v ′k ∈ Tp′ (P ) are such that P∗p′ (v ′i ) = w i for i = 1, . . . , k.
We must show that
φp′ (v ′1 , . . . , v ′k ) = φp (v 1 , . . . , v k ). (4.5.3)
where
v ′′i = (σg )−1 ′ ′
∗p (v i ) = (σg −1 )∗p·g (v i ), i = 1, . . . , k
(by assumption #2).
Exercise 4.5.2 Show that P∗p (v ′′i ) = w i for i = 1, . . . , k.
Thus, P∗p (v i − v ′′i ) = 0 so v i − v ′′i is vertical for i = 1, . . . , k. By assumption
#1, φp (v 1 − v ′′1 , v 2 , . . . , v k ) = 0 so
φp (v 1 , v 2 , . . . , v k ) = φp (v ′′1 , v 2 , . . . , v k ) .
φp (v 1 , v 2 , v 3 , . . . , v k ) = φp (v ′′1 , v ′′2 , v 3 , . . . , v k ).
We will have quite a bit more to say about projectable forms on bundles,
but first we derive some general results on derivatives of tensorial forms. First
we show that the exterior derivative of a pseudotensorial form is itself pseu-
dotensorial.
P
Lemma 4.5.2 Let G ,→ P −→ X be a smooth principal bundle, V a finite
dimensional real vector space, ρ : G −→ GL(V) a representation of G on V
and φ a V-valued k-form on P that is pseudotensorial of type ρ. Then dφ is
pseudotensorial of type ρ.
P
Theorem 4.5.4 Let G ,→ P −→ X be a smooth principal bundle with a con-
nection ω, V a finite dimensional real vector space, ρ : G −→ GL(V) a repre-
sentation of G on V and φ a V-valued k-form on P that is pseu-dotensorial
of type ρ. Then the covariant exterior derivative dω φ of φ, defined by
( )
(dω φ)(p)(v1 , . . . , vk+1 ) = (dφ)H (p) v1 , . . . , vk+1
( )
= dφ(p) vH1 , . . . , vH
k+1
( )
= (P ∗ (dφ̄))p v H H
1 , . . . , v k+1
( )
= (d(P ∗ φ̄))p v H H
1 , . . . , v k+1
( )
= (dφ)p v H H
1 , . . . , v k+1
d d
A·v = (exp(tA) · v)|t=0 = (ρ(exp(tA))(v))|t=0 ,
dt dt
where ρ : G −→ GL(V) is the representation relative to which our forms
are tensorial. Notice that if ρ is just the natural representation of G ⊆
GL(k, C) on Ck (matrix multiplication), then A · v = Av is also matrix mul-
tiplication for each A ∈ G (Exercise 6.8.6, [N4]). A special case of more
immediate concern to us here is the following: Suppose ρ = ad : G −→ GL(G)
is the adjoint representation. Then, for any A, B ∈ G,
4.5. Tensorial Forms 227
d( )
A·B = adexp(tA) (B) |t=0
dt
1( )
= lim adexp(tA) (B) − B
t−→0 t
= [A, B],
where this last equality follows from the definition of the Lie bracket and is
proved on pages 288–289 of [N4].
Now, the bilinear map (A, v) −→ A·v of G×V to V determines a wedge prod-
uct for G-valued forms and V-valued forms. Specifically, if α ∈ Λk (P, G) and
β ∈ Λl (P, V), then we define α∧β
˙ ∈ Λk+l (P, V) at each point of P by
( ) 1 ∑ ( )
(α∧β)
˙ v 1 , . . . , v k+l = (−1)σ α v σ(1) , . . . , v σ(k)
k! l! σ
( )
· β v σ(k+1) , . . . , v σ(k+l) ,
where the sum is over all permutations σ ∈ Sk+l of {1, . . . , k + l}. Notice, in
particular, that when ρ is the adjoint representation of G on G, then α∧β ˙ is
just the bracket wedge product [α, β] described in Section 4.2. We are now in
a position to prove our major result.
P
Theorem 4.5.6 Let G ,→ P −→ X be a smooth principal bundle with a con-
nection ω, V a finite dimensional real vector space, ρ : G −→ GL(V) a repre-
sentation of G on V and φ ∈ Λkρ (P, V) a V-valued, tensorial k-form of type ρ.
Then
dω φ = dφ + ω ∧φ.˙
1 ∑ ( )
+ (−1)σ ω p v σ(1) (4.5.5)
k!
σ∈Sk+1
( )
· φp v σ(2) , . . . , v σ(k+1) .
From Exercise 4.5.5 and the fact that V 1 is vertical and φ is horizontal we
conclude that
( )
d φ(V 1 , V 2 , . . . ,V k+1 ) = (−1)1+1 V 1 φ (V 2 , . . . ,V k+1 ) + 0
( )
= V 1 φ (V 2 , . . . ,V k+1 ) .
230 4. Differential Forms and Integration
Since the left-hand side of (4.5.5) is obviously zero in Case III, the proof now
reduces to showing that
( ) 1 ∑ ( ) ( )
0 =V 1 φ (V 2 , . . . ,V k+1 ) + (−1)σ ω Vσ(1) · φ Vσ(2) , . . . ,Vσ(k+1) .
k!
σ ∈ Sk+1
But since ω(V σ(1) ) will be zero whenever σ(1) ̸= 1 we may rewrite this as
( ) 1 ∑ ( )
0 = V 1 φ (V 2 , . . . ,V k+1 ) + (−1)σ ω(V1 ) · φ V σ(2) , . . . ,Vσ(k+1)
k! σ∈S
k+1
σ(1)=1
( )
1 ∑ ( )
0 = V 1 φ (V 2 , . . . ,V k+1 ) +ω(V1 ) · (−1)σ φ Vσ(2) , . . . ,Vσ(k+1)
k! σ∈Sk+1
σ(1)=1
( )
0 = V 1 φ (V 2 , . . . ,V k+1 ) + ω(V 1 ) · φ(V 2 , . . . ,V k+1 ). (4.5.6)
Thus,
( ) d( )
V 1 (p) φ (V 2 , . . . ,V k+1) = exp (t(−A)) · φp (V 2 (p), . . . ,Vk+1 (p))
dt t=0
( )
= (−A) · φp V 2 (p), . . . ,V k+1 (p)
( )
= −ω p (A# ) · φp V 2 (p), . . . ,V k+1 (p)
( )
= −ω p (V 1 (p)) · φp V 2 (p), . . . ,V k+1 (p)
which gives (4.5.6) at the arbitrary point p ∈ P and therefore completes the
proof.
P
Corollary 4.5.8 Let G ,→ P −→ X be a smooth principal bundle with a
connection ω and ad : G −→ GL(G) the adjoint representation of G on G.
Then, for each φ ∈ Λkad (P, G),
d ω φ = d φ + [ω, φ].
Notice that the Corollary does not apply to ω itself, which is pseudotensorial
of type ad, but not horizontal. Indeed, the Cartan Structure Equation gives
d ω ω = Ω = d ω+ 12 [ω, ω]. However, the Corollary does apply to the curvature
Ω and gives d ω Ω = d Ω + [ω, Ω] which we now show is identically zero.
P
Theorem 4.5.9 (Bianchi Identity) Let G ,→ P −→ X be a smooth prin-
cipal bundle with connection ω and curvature Ω = d ω ω. Then
d ω Ω = 0. (4.5.7)
Proof: We simply compute, from Corollary 4.5.8,
d ω Ω = d Ω + [ω, Ω]
( ) [ ]
1 1
= d d ω + [ω, ω] + ω, d ω + [ω, ω]
2 2
1 1
= d (d ω) + d ([ω, ω]) + [ω, d ω] + [ω, [ω, ω]]
2 2
1
= d ([ω, ω]) + [ω, d ω]
2
1
= ([d ω, ω] − [ω, d ω]) + [ω, d ω]
2
1
= (−[ω, d ω] − [ω, d ω]) + [ω, d ω] = 0.
2
Since [ω, Ω] = −[Ω, ω] one can write (4.5.7) as
d Ω = [Ω, ω]. (4.5.8)
232 4. Differential Forms and Integration
∫ ∫ π ( √ )
4
f= sec = ln|sec + tan| = ln 1 + 2 .
[0,1] [0, π 0
4]
π
4
= ln|sec u + tan u|
0
( √ )
= ln 1 + 2 .
There is much to be said for this. The student’s calculations, which, on the
surface, appear more formal and less rigorous, are, in fact, entirely rigorous
once one has faced the fact that it is not functions, but 1-forms that should
be integrated over intervals in R. Specifically, any f : R −→ R gives rise
to a unique 1-form ω = f dx on R and, if g is an orientation preserving
diffeomorphism with g([α, β]) = [a, b], then
φ−1 ([α1 , β1 ]) = [a, b] and ψ −1 ([α2 , β2 ]) = [a, b], then our claim is that
∫ ∫
h ◦ φ−1 = k ◦ ψ −1
[α1 ,β1 ] [α2 ,β2 ]
∫
so that [a,b] ω could be defined as the integral of any of its coordinate ex-
pressions. To prove this we define
g = ψ ◦ φ−1 : φ (U ∩ V ) −→ ψ (U ∩ V ).
(k ◦ ψ −1 ) ◦ g = k ◦ φ−1 .
as required.
This we regard as rather persuasive evidence in favor of the view that it
is not functions, but 1-forms that one should integrate on 1-manifolds. There
is no standard coordinate system on a general 1-manifold and therefore no
standard coordinate expression for a function on a 1-manifold and, even in
R, it matters very much which coordinate expression one chooses to integrate
over a given subset. Coordinate expressions for 1-forms, on the other hand,
have just what it takes (according to (4.6.1)) to possess invariant integrals.
That all of this works out just as nicely for n-forms on n-manifolds will become
clear from a glance at Exercise 4.3.4 and the Change of Variables Formula for
Rn (sans a few hypotheses that we will add a bit later): Suppose U and V are
open in Rn and g : U −→ V is a diffeomorphism. If M ⊆ U and f : V −→ R,
then ∫ ∫
f= (f ◦ g) det(g ′ ) ,
g(M ) M
where g ′ is the Jacobian of g, i.e., the matrix of g∗ relative to standard coor-
dinates.
With this motivation behind us we set about building a theory of integration
for n-forms on n-dimensional manifolds. We will presume a familiarity with
the basics of Lebesgue integration on Rn .
4.6. Integration on Manifolds 235
Remark: This decision to build upon the Lebesgue rather than the Riemann
integral smooths the theoretical development in the early stages, but has es-
sentially no practical impact on the explicit calculations we must perform. The
reader not versed in the Lebesgue theory has a number of options. Perhaps the
most sensible is to simply ignore all references to the subject and be assured
that, in the end, the integrals we must actually compute are accessible to the
tools available in the theory of the Riemann integral. Alternatively, one might
actually learn something about this indispensible part of modern mathemat-
ics. We will provide a brief synopsis, but everything we need (and more) is
covered very nicely and in fewer than fifty pages (specifically, pages 201–247)
in [Brow].
Throughout the remainder of this section X will denote a smooth, oriented,
n-dimensional manifold with n ≥ 1 and all charts will be assumed consistent
with the given orientation.
Recall that a subset A of Rn is said to have (Lebesgue) measure zero
in Rn if it can be covered by countable families of open rectangles with ar-
bitrarily small total volume. In more detail, an open rectangle in Rn is a set
of the form R = (a1 , b1 ) × · · · × (an , bn ), where ai < bi for i = 1, . . . , n, and
its volume is vol(R) = (b1 − a1 ) · · · (bn − an ). Then A ⊆ Rn has measure
zero if, for∪every ϵ > 0,∑there exists a family R1 , R2 , . . . of open rectangles
∞ ∞
with A ⊆ i=1 Ri and i=1 vol(Ri ) < ϵ. Subsets of sets of measure zero ob-
viously also have measure zero and countable unions of sets of measure zero
also have measure zero (cover the k th set with a family of rectangles of total
volume less than ϵ/2k ). It is also true, but not at all obvious, that smooth
images of sets of measure zero have measure zero, i.e., if U ⊆ Rn is open,
f : U −→ Rn is smooth and A ⊆ U has measure zero, then f (A) has measure
zero (Proposition 5–17 of [N1], or Lemma 1.1, Chapter 3, of [Hir]).
A subset A of a smooth manifold X is said to have measure zero in X if,
for every chart (U, φ) for X, the set φ(U ∩ A) has measure zero in Rn .
Exercise 4.6.1 Show that A has measure zero in X if and only if, for each
a ∈ A, there exists a chart (U, φ) at a for which φ(U ∩ A) has measure zero in
Rn . Hint: Keep in mind that our manifolds are assumed second countable.
If S is any set, then a collection A of subsets of S is called a σ-algebra if it
contains the empty set and is closed under the formation of complements and
countable
∪∞ unions, i.e., if (1) ∅ ∈ A, (2) S − A ∈ A whenever A ∈ A, and (3)
i=1 A i ∈ A whenever Ai ∈ A for i = 1, 2, . . . . Any collection C of subsets
of S is contained in various σ-algebras of subsets of S (e.g., the collection of
all subsets of S) and the intersection of all of these σ-algebras containing C is
itself a σ-algebra. This intersection is called the σ-algebra generated by C. In
any topological space S the σ-algebra generated by the collection C of closed
sets is called the Borel σ-algebra and its elements are called Borel sets in
S. Closed sets, open sets, countable unions of countable intersections of open
sets, etc., are all Borel sets.
236 4. Differential Forms and Integration
The collection of Borel sets in Rn is huge, but we need to enlarge the col-
lection still further. A subset M of Rn is said to be (Lebesgue) measurable
if it can be written as M = A ∪ B, where A ∩ B = ∅, A has measure zero in
Rn and B is a Borel set in Rn . Thus, measurable sets are those which differ
from a Borel set by a set of measure zero. Since the empty set has measure
zero, every Borel set is measurable. Since the empty set is a Borel set, every
set of measure zero is measurable.
Remark: Those of our readers who are “in the know” realize that this is
not the usual definition of Lebesgue measurable. That it is equivalent to the
usual definition follows from Theorem 9.29 of [Brow].
The collection M of Lebesgue measurable sets is itself a σ-algebra.
As the name suggests, the Lebesgue theory assigns to every M ∈ M a
“measure” m(M ) of its size. For example, m(R) = vol(R) for every rectan-
gle R and m(A) = 0 for any set A of measure zero. The ability to measure a
large collection of sets provided Lebesgue with the means to integrate a large
collection of functions. Roughly, his idea was as follows: Riemann integrated
a function f : [a, b] −→ R by partitioning [a, b] into subintervals, approximat-
ing f by a step function that was constant on each subinterval, defining the
integral of the step function in the only reasonable way and taking the limit
as the partition became finer and finer.
a b
This limit will not exist unless f is relatively nice, e.g., a bounded, real-valued
function on [a, b] must be continuous except perhaps on a set of measure zero
(Theorem 10.23 of [Brow]). This is unfortunate since, for example, limits of
even very nice functions are often not at all nice so the Riemann integral
does not react well to taking limits. Lebesgue’s idea was to partition, not the
4.6. Integration on Manifolds 237
ci
Mi
For the record we will now briefly sketch the construction of the Lebesgue in-
tegral in somewhat more detail and formulate those definitions and results from
the theory that we will need. A real-valued function f on Rn is measurable
if the inverse image of every closed interval in R is Lebesgue measurable in
Rn (if f is an extended real-valued function, i.e., maps to R ∪ {±∞}, then it
is also required that f −1 (∞) and f −1 (−∞) be measurable). A (measurable)
step function is a function s that is zero except on a finite number of disjoint
measurable sets M1 , . . . , Mk with m(Mi ) < ∞ for each i = 1, . . . , k and which
takes a finite constant value ci on Mi for i = 1, . . . , k. The integral of such a
step function s over Rn is defined by
∫ ∑k
s dm = ci m(Mi ).
Rn i=1
Sets of measure zero “don’t count,” i.e., if f is integrable and g agrees with
f except perhaps on a set of measure zero (we say that g equals f almost
everywhere and write g = f a.e.), then g is integrable and
∫ ∫
g dm = f dm (g = f a.e.).
M M
(Theorem 10.46 of [Brow]). A very useful result for the purposes of calculation
is that if f is continuous on a closed∫ rectangle R̄ = [a1 , b1 ] × · · · × [an , bn ], then
it is Lebesgue integrable on R̄ and R̄ f dm agrees with the ordinary Riemann
integral of f over R̄. Consequently, the value of the integral can be calculated
as an iterated integral (Fubini’s Theorem) from the Fundamental Theorem of
Calculus:
∫ ∫ 1 (∫ 2 (
b b ∫ n b
) )
f dm = ··· f (x1 , x2 , . . . , xn ) dxn · · · dx2 dx1 .
R̄ a1 a2 an
Exercise 4.6.2 Show that M ⊆ X is measurable if and only if, for each
m ∈ M , there exists a chart (U, φ) at m for which φ(U ∩ M ) is Lebesgue
measurable in Rn . Hint: Use Exercise 4.6.1 and the fact that diffeomorphisms
on open sets in Rn preserve measure zero, Borel sets and disjoint unions.
240 4. Differential Forms and Integration
so ( )
k ◦ ψ −1 = (hα ◦ φ−1 ′
α ) ◦ g det(g ) (4.6.5)
on ψ(Uα ∩ V ). Now, on open sets in R , smooth maps are measurable and
n
We define the integral of ω over X first in the case in which supp ω is compact
and contained in some coordinate neighborhood and then, in the general case,
use a partition of unity to define the integral as a sum of integrals of the first
type.
Suppose then that ω is a measurable n-form on X with compact support
supp ω ⊆ U , where (U, φ) is a chart on X (keep in mind that we consider
only charts consistent with the given orientation of X). If x1 , . . . , xn are the
coordinate functions of φ and ω = h dx1 ∧ · · · ∧ dxn on U , then the coordinate
expression h ◦ φ−1 is measurable on the open set φ(U ) in Rn . We say that
4.6. Integration on Manifolds 241
Now, since both (U, φ) and (V, ψ) are consistent with the orientation of X,
the Jacobian determinant det(g ′ ) is positive so we may, if we wish, write this
as
k ◦ ψ −1 = ((h ◦ φ−1 ) ◦ g) det(g ′ ) .
But now the Change of Variables Formula (4.6.3) with f = h ◦ φ−1 and
M = ψ(U ∩ V ) assures the integrability of k ◦ ψ −1 and gives
∫ ∫
h ◦ φ−1 dm = ((h ◦ φ−1 ) ◦ g) det(g ′ ) dm
φ(U ∩V ) ψ(U ∩V )
∫
= k ◦ ψ −1 dm
ψ(U ∩V )
as required.
Before proceeding with the general case we make a few observations about
this one. First notice that if ω 1 and ω 2 are two n-forms of the prescribed type
with supports both contained in U , then ω 1 + ω 2 is also of this type and
∫ ∫ ∫
ω1 + ω2 = ω1 + ω2 .
X X X
242 4. Differential Forms and Integration
Now let us turn to the general case and consider a measurable n-form ω
on X(supp ω need not be compact and need not be contained in a coordi-
nate neighborhood). Choose an oriented atlas for X and a partition of unity
{ϕk }k=1,2,... subordinate to it. Then each ϕk ω is an n-form with compact
support contained in ∑ a coordinate neighborhood. Since {supp ϕk }k=1,2,... is
∞
locally finite the sum k=1 ϕk ω is finite on some neighborhood of each point
so
∑∞
ω = ϕk ω
k=1
Similarly,
∫ ∞ ∫
∑
ϕk ω = ϕ′j ϕk ω.
X j=1 X
Thus,
∞ ∫
∑ ∞ ∫
∞ ∑
∑
ϕ′j ω= ϕk ϕ′j ω
j=1 X j=1 k=1 X
and ∞ ∫
∑ ∞ ∫
∞ ∑
∑
ϕk ω = ϕ′j ϕk ω. (4.6.8)
k=1 X k=1 j=1 X
∑∞ ∫
Exercise 4.6.4 Show that absolute convergence of k=1 X ϕk ω implies the
convergence of
∑∞ ∑∞ ∫
ϕ′j ϕk ω
k=1 j=1 X
∑∞ ∫
Finally, conclude that j=1 X ϕ′j ω converges absolutely.
Remark: If ω has compact support (not necessarily contained in a coordinate
neighborhood), then all of the sums dealt with above are finite so absolute con-
vergence is assured. In particular, this is always the case when X is compact.
To complete our sequence of definitions we let ω be a measurable n-form
on X and M ⊆ X a measurable set. We say that ω is integrable on M if
χM ω is integrable on X and in this case, we define
∫ ∫
ω= χM ω.
M X
where the convergence on the right-hand side is absolute. We must show that
∞ ∫
∑ ∞ ∫
∑
∗
(ϕk ◦ f )f ω = ϕk ω
k=1 X k=1 Y
and that the convergence on the left-hand side is absolute. Both of these will
be proved if we show that
∫ ∫
∗
(ϕk ◦ f )f ω = ϕk ω
X Y
∗ ∗
for each k. Notice that (ϕk ◦ f )f ω = f (ϕk ω) so we must prove that
∫ ∫
f ∗ (ϕk ω) = ϕk ω.
X Y
∫ ∫
−1
= h◦f ◦f ◦ φ−1
α dm = h ◦ φ−1
α dm
φα (Uα ) φα (Uα )
∫
= ϕk ω
Y
as required.
Exercise 4.6.6 Show that, if f : X −→ Y is an orientation reversing diffeo-
morphism, then ∫ ∫
f ∗ω = − ω.
X Y
It’s time now to actually compute a few integrals. We recall (Theorem 4.3.2)
that any oriented Riemannian n-manifold X has defined on it a unique
246 4. Differential Forms and Integration
Now,
(ι ◦ f )∗ (x1 dx2 − x2 dx1 ) = (x1 ◦ ι ◦ f ) d(x2 ◦ ι ◦ f )
− (x2 ◦ ι ◦ f ) d(x1 ◦ ι ◦ f )
= cos θ d(sin θ) − sin θ d(cos θ)
= cos2 θ dθ + sin2 θ dθ
= dθ
so ∫ ∫
ω = dθ.
S1 (−π,π)
Thus, ∫ ∫
(
ω = (ι ◦ f )∗ x1 dx2 ∧ dx3
S2 (0,π)×(−π,π)
)
− x dx ∧ dx3 + x3 dx1 ∧ dx2 .
2 1
Now,
(ι ◦ f )∗ (x1 dx2 ∧ dx3 ) = sin ϕ cos θ (d(sin ϕ sin θ) ∧ d(cos ϕ))
= sin ϕ cos θ ((cos ϕ sin θ dϕ + sin ϕ cos θ dθ)
∧ (− sin ϕ dθ))
= sin3 ϕ cos2 θ dϕ ∧ dθ.
Note: It will follow from Stokes’ Theorem (Section 4.7) that the integral of
ω over a circle in R2 − {(0, 0)} is 2π if the circle encloses the origin and 0 if
it does not.
We conclude this section with a calculation that yields a special case of
Stokes’ Theorem.
Theorem 4.6.2 Let X be a smooth, oriented, n-dimensional manifold and
suppose ω ∈ Λn−1 (X) is a smooth (n − 1)-form on X with compact support.
Then ∫
dω = 0.
X
4.7. Stokes’ Theorem 249
With a glace back at Exercise 4.4.8 it may come as no surprise that these
theorems are really statements about (n − 1)-forms and their exterior deriva-
tives. Remarkably, these are all, in a sense, the same statement and it is our
objective here to derive the one simple, elegant equality from which they all
follow: ∫ ∫
∗
ι ω= dω. (4.7.1)
∂D D
The first order of business is to define precisely the appropriate regions of
integration.
Throughout this section X will denote a smooth, n-dimensional, oriented
manifold. A subset D of X will be called a domain with smooth boundary
in X if, for each p ∈ X, one of the following is true:
(1) There is an open neighborhood of p in X which is contained entirely in
X − D (such points are said to be in the exterior of D).
(2) There is an open neighborhood of p in X which is contained entirely in
D (these are called the interior points of D).
(3) There exists a chart (U, φ) for X at p with φ(p) = 0 ∈ Rn and φ(U ∩
D) = φ(U ) ∩ Rn+ , where Rn+ = {(x1 , . . . , xn ) ∈ Rn : xn ≥ 0} is the
closed upper half-space in Rn (these are called the boundary points
of D and the set of all such is denoted ∂D and called the boundary
of D).
Remark: These definitions make sense for any n ≥ 1, but, when n = 1,
∂D is a set of isolated points. Orientations for 0-dimensional manifolds and
integrals over such can all be defined in such a way that the results of this
section remain valid in this case, but we will have no need of them and so will
assume henceforth that n ≥ 2. Notice also that the set Int D of all interior
points of D is open in X and that there is nothing in the definition to preclude
the possibility that ∂D = ∅ (so X itself qualifies as a domain with smooth
boundary).
Exercise 4.7.1 Show that these three conditions are mutually exclusive and
that, for any chart (U, φ) of the type described in (3), φ(U ∩ ∂D) = {x ∈
φ (U ) : xn = 0}. Conclude (e.g., from Exercise 5.6.1, [N4]) that ∂D is an
(n − 1)-dimensional submanifold of X.
Exercise 4.7.2 Show that the n-dimensional disc Dn = {x ∈ Rn : ∥x∥ ≤ 1}
is a domain with smooth boundary in Rn and ∂Dn = S n−1 .
Exercise 4.7.3 Show that if D is a domain with smooth boundary in X,
then ∂D has measure zero in X and conclude that D is measurable in X.
Next we wish to show that, when D is a domain with smooth boundary in
X, the submanifold ∂D inherits a natural orientation from X.
Remark: This induced orientation on ∂D is “natural” in the somewhat
indirect sense that it supplies ∂D with the orientation it must have in order
4.7. Stokes’ Theorem 251
∂ ∂
, . . . , n−1 .
∂x1 ∂x
p p
X = Rn ).
Exercise 4.7.4 Show that this definition does not depend on the choicϵ of
(U, φ), i.e., that if (V, ψ) is another chart at p in X, consistent with the
orientation of X, and with ψ(p) = 0 and ψ(V ∩ D) = ψ(V ) ∩ Rn+ and u =
i |p , where y , . . . , y
∂
bi ∂y 1 n−1
, y n are the coordinate functions of ψ, then an and
n
b have the same sign.
Now we define the induced orientation for Tp (∂D) by decreeing that a basis
{v 1 , . . . , v n−1 } for Tp (∂D) is in this orientation if and only if {v , v 1 , . . . , v n−1 }
is an oriented basis for Tp (X) for any outward pointing tangent vector
u ∈ Tp (X).
Exercise 4.7.5 Show that the induced orientation for Tp (∂D) is well-defined,
i.e., does not depend on the choice of outward pointing vector u ∈ Tp (X).
This sum is locally finite and therefore defines a smooth vector field on U ⊇ ∂D
which is outward pointing on ∂D since 0 ≤ ϕn ≤ 1.
Now, let ω be a nonzero n-form on X that determines the orientation of
X (i.e., ω p (v 1 , . . . , v n−1 , v n ) > 0 whenever p ∈ X and { v 1 , . . . , v n−1 , v n }
is an oriented basis for Tp (X)). Regard ω as a C ∞ (X)-multilinear map on
n
X (X) × · · · × X (X). The restriction (pullback) of ω to U , which we will
n
also denote ω, operates on X (U ) × · · · × X (U ). Thus, we may define an
(n − 1)-form ω̃ on U by
ω̃ (V 1 , . . . ,V n−1 ) = ω (V , V 1 , . . . ,V n−1 ),
Proof: Observe first that, since ω is smooth with compact support, the
same is true of ι∗ ω and d ω so both of these are integrable. Now, choose a
countable, oriented atlas { (Uk , φk ) }k=1,2,... for X such that each φk (Uk ) is a
bounded subset of Rn and either Uk ∩∂D = ∅ (if Uk ∩D ̸= ∅ we assume in this
case that Uk ⊆ Int D), or, if Uk ∩∂D ̸= ∅, then φk (Uk ∩D) = φk (Uk )∩ Rn+ . Let
{ ϕk }k=1,2,... be a family of real-valued
∑∞ functions on X of the sort guaranteed
by Corollary 3.1.5. Then ω = ∑∞
k=1 ϕk ω, where the sum has ∑only finitely
∞
many nonzero terms. Thus d ω = k=1 d(ϕk ω) and ι∗ ω = k=1 ι∗ (ϕk ω)
are finite sums, as are each of the following:
∫ ∞ ∫
∑
∗
ι ω= ι∗ (ϕk ω)
∂D k=1 ∂D
∫ ∞ ∫
∑
dω = d(ϕk ω).
D k=1 D
for each k and this simply amounts to proving (4.7.2) in the special case in
which supp ω is contained in some Uk .
Assume then ∫that supp ω ⊆ Uk and consider first the
∫ case in which Uk ∩
∂D = ∅. Then ∂D ι∗ ω = 0 and we must show that D d ω is zero as well.
This is obvious if Uk ∩ D = ∅ since supp dω ⊆ supp ω ⊆ Uk . If Uk ∩ D ̸= ∅,
then Uk ⊆ Int D so
∫ ∫ ∫
dω = dω = dω
D Uk X
on Uk (including the (−1)i−1 here will spare us a few of these down the road).
Thus,
( n )
∑ ∂
−1
dω = i
(ωi ◦ φk ) ◦ φk dx1 ∧ · · · ∧ dxn (4.7.3)
i=1
∂x
Exercise 4.7.9 Argue as in the proof of Theorem 4.6.2 to show that, for
i = 1, . . . , n − 1 (but not n),
∫
∂(ωi ◦ φ−1
k )
χRn+ i
dm = 0 (4.7.8)
C ∂x
4.7. Stokes’ Theorem 255
so
∫ ∫ ∫
dω = − (ωn ◦ φ−1 1
k ) (x , . . . , x
n−1
, 0) dx1 · · · dxn−1 = ι∗ ω
D R n−1 ∂D
as required.
We conclude with an analogue of the familar principle of deformation of
paths for line integrals in vector calculus. Let X be a smooth, oriented
n-dimensional manifold and let N1 and N2 be two oriented k-dimensional
submanifolds, where 1 ≤ k ≤ n. We say that N1 can be smoothly deformed
into N2 if there exists a smooth map H : N × (0, 3) −→ X such that
H|N × {1} −→ N1 and H|N × {2} −→ N2 are orientation preserving diffeo-
morphisms. We claim that, if ω is any closed k-form on X and ι1 : N1 ,→ X
and ι2 : N2 ,→ X are the inclusion maps, then
∫ ∫
ι∗1 ω = ι∗2 ω . (4.7.10)
N1 N2
Introduction
The plane R2 and the punctured plane R2 − {(0, 0)} are not diffeomorphic,
nor even homeomorphic. There are various means by which one can prove
this, but the most instructive among these detect the “hole” in the punctured
plane (and none in R2 ) by distinguishing topologically certain “types” of cir-
cles that can live in the two spaces. One way of formalizing this idea is to show
that R2 and R2 − { (0, 0) } have different fundamental groups (see pages 107
and 133 of [N4]). Fundamental groups (and their higher dimensional gener-
alizations) are notoriously difficult to calculate, however. In this chapter we
will investigate another method of attaching to any smooth manifold certain
algebraic objects which likewise “detect holes”. The idea is quite simple and,
for R2 and R2 − { (0, 0) }, has already been hinted at in Section 4.4. There we
found that any closed 1-form on R2 is necessarily exact (Poincaré Lemma),
but constructed a closed 1-form ω on R2 − { (0, 0) } that is not exact. Stokes’
Theorem implies that the integral of any 1-form on R2 over any circle (com-
pact, connected, 1-dimensional submanifold) is zero and the same is true of
ω provided the circle does not contain (0, 0) in its interior. For a circle S 1
that does contain (0, 0) in its ∫interior (and has the orientation induced from
R2 − { (0, 0) }) we computed S 1 ω and obtained 2π. The idea then is that
the closed, nonexact 1-form ω detects the hole in R2 − { (0, 0) } via its inte-
grals over circles. The prospect then arises that one might distill topological
information from the existence of closed, nonexact forms on a manifold. Un-
fortunately, the collection of all such forms (even of some fixed degree) can be
a huge, unmanageable object. Observe, however, that another closed 1-form
ω ′ on R2 − { (0, 0) } that differs from ω by an exact form (ω ′ = ω + dη)
has, by Stokes’ Theorem, the same integral as ω over any circle and so for
our purposes, there is no reason to distinguish them. The natural thing to do
then is define an equivalence relation on the set of closed 1-forms whereby two
such forms are equivalent (we shall say “cohomologous”) if they differ by an
exact form. The resulting set of equivalence classes inherits a natural vector
space structure from the space of 1-forms and, with this structure, is called
the 1st de Rham cohomology group of the manifold. For R2 it has dimension 0
(no holes), while for R2 − { (0, 0) } it has dimension 1 (one hole). The precise
construction of this vector space and its higher dimensional analogues is the
task we set for ourselves in the next section. There are many formal similari-
ties with homology theory (Chapter 3 of [N4]), but the arguments tend to be
substantially easier.
d d d
−→ Λn−1 (X) −→ Λn (X) −→ · · ·
The efficient calculation of cohomology groups will require that we first deal
with some of the more formal aspects of the subject. There is, however, at least
one small example accessible to us at this stage. We know that Hde0 R (S 1 ) = R
and Hdek R (S 1 ) = 0 for k ≥ 2 and k ≤ −1. Furthermore, Hde1 R (S 1 ) has
dimension at least one by Lemma 5.1.2. We show that, in fact,
Hde1 R (S 1 ) ∼
= R. (5.1.1)
260 5. de Rham Cohomology
Since we know from Lemma 5.1.2 that dim Hde1 R (S 1 ) ≥ 1 we need only prove
that dim Hde1 R (S 1 ) ≤ 1 in order to establish (5.1.1). For this it will suf-
fice to find a nontrivial cohomology class [η] ∈ Hde1 R (S 1 ) such that, for
any 1-form α on S 1 , there exists a constant α for which α − αη is exact
(since then [α] = [αη] = α[η]). For this we let ι : S 1 ,→ R2 − { (0, 0) }
be the inclusion of S 1 into the punctured plane and define η = ι∗ ω, where
ω is the 1-form on R2 − { (0, 0) } defined (in Section 4.4) from two angular
coordinate functions on R2 − { (0, 0) }. Specifically,
y x
ω=− dx + 2 dy
x2 +y 2 x + y2
1
so
∫ η is just the standard volume form1 on S1 . In Section 4.6 we found that
S1
η = 2π so, in particular, [η] ∈ Hde R (S ) is nontrivial. Now let α be an
arbitrary 1-form on S 1 and set
∫
1
α= α.
2π S 1
[ω 1 ] ∧ [ω 2 ] = [ω 1 ∧ ω 2 ]
262 5. de Rham Cohomology
(g ◦ f )# = f # ◦ g # . (5.2.3)
hk ƒ* – g*
hk + 1
∑
Note that P sends the term J ωJ (x, t)dxJ to zero and that P ω has no dt
(although its components depend on t). Since these components are clearly
C ∞ , P ω is indeed a (k − 1)-form.
Now, suppose V is another open set in Rn and ϕ : V → U is a diffeomor-
phism. Then
Φ = ϕ × idR : V × R −→ U × R
is a diffeomorphism and, for each k = 0, 1, . . .
Φ∗ : Λk (U × R) −→ Λk (V × R).
We claim that
Φ∗ (P ω) = P (Φ∗ ω) (5.2.6)
for each ω ∈ Λ (U × R).
k
5.2. Induced Homomorphisms 265
P
Λk ( U × ) Λk – 1 ( U × )
Φ* Φ*
Λk ( V × ) Λk – 1 ( V × )
P
where K varies over increasing index sets of length k and the functions γK do
not depend on t. An analogous statement is true of each Φ∗ (dxJ ). Now, for
ω ∈ Λk (U × R),
( )
∑ ∑
Φ∗ ω(x, t) = Φ∗ ωI (x, t) dt ∧ dxI + ωJ (x, t) dxJ
I J
∑ ( ) ∑ ( )
∗
= ωI (x, t)Φ dt ∧ dx I
+ ωJ (x, t) Φ∗ dxJ
I J
∑ ( ) ∑ ( )
= ωI (x, t) dt ∧ Φ∗ dxI + ωJ (x, t) Φ∗ dxJ .
I J
266 5. de Rham Cohomology
The fact that P , defined for open subsets of Rn , commutes with pullbacks
by diffeomorphisms of the type Φ = ϕ × idR permits us to define an analogous
operation on manifolds. More precisely, we will prove the following: Let X be
a smooth n-manifold. Then, for each k = 0, 1, . . . there exists a linear map
P : Λk (X × R) −→ Λk−1 (X × R)
such that
1. If X = U , an open submanifold of Rn , then P is given by (5.2.5).
2. If Y is another smooth n-manifold, ϕ : Y → X is a diffeomorphism and
Φ = ϕ × idR : Y × R → X × R, then
Φ∗ ◦ P = P ◦ Φ∗ . (5.2.7)
To prove this we first take P to be identically zero when k = 0. Now let
ω ∈ Λk (X × R), where k ≥ 1. We will define P ω locally in coordinates and
then show that the definition is independent of the choice of coordinates. Let
(U, φ) be a chart on X. Then φ−1 : φ(U ) → U is a diffeomorphism and
therefore so is φ−1 × idR : φ(U ) × R → U × R. Thus, (φ−1 × idR )∗ ω is a
k-form on φ(U ) × R ⊆ Rn × R (technically we should write ι∗ ω rather than
ω, where ι : U × R ,→ X × R, but we will suppress this inclusion). Thus,
P ((φ−1 × idR )∗ ω) as defined above in Euclidean spaces, is a (k − 1)-form on
φ(U ) × R. We may therefore define P ω on U by
( ( )
( )∗ )
P ω = (φ × idR )∗ P φ−1 × idR ω .
Thus,
( (( )∗ ))
∗ −1
(ψ × idR ) P ψ × idR ω
(( ( )
))
∗ ∗ −1 ∗
= (ψ × idR ) Φ P (φ × idR ) ω
( )∗ ( (( )∗ )
)
−1
= Φ ◦ (ψ × idR ) P φ × idR ω
( ( )
∗
( −1 )∗ )
= (φ × idR ) P φ × idR ω
π ia
U× U U×
Φ Φ
π ia
268 5. de Rham Cohomology
π* i*a
Λk ( U × ) Λk ( U) Λk ( U × )
Φ* Φ*
π* ia*
i.e., ( )
(ia ◦ π)∗ (Φ∗ ω̃) = Φ∗ (ĩa ◦ π̃)∗ ω̃ .
Moreover,
ω − (ia ◦ π)∗ ω = Φ∗ (ω̃) − (ia ◦ π)∗ (Φ∗ ω̃)
= Φ∗ (ω̃ − (ĩa ◦ π̃)∗ ω̃).
Thus, assuming we have proved that d ◦ P (ω̃) + P ◦ d(ω̃) = ω̃ − (ĩa ◦ π̃)∗ ω̃,
the result for X will follow on U and therefore on all of X.
Thus, we need only prove the result when X = U is an open subman-
ifold of Rn (we will drop all of the tildas in the notation used above).
Any ω ∈ Λk (U × R) has a unique representation in standard coordinates
(x, t) = (x1 , . . . , xn , t) of the form
∑ ∑
ω(x, t) = ωI (x, t) dt ∧ dxI + ωJ (x, t) dxJ .
I J
But
(ia ◦ π)∗ (ωJ (x, t)) = ωJ ◦ (ia ◦ π)(x, t) = ωJ (x, a)
and
(ia ◦ π)∗ (dxj ) = d (xj ◦ (ia ◦ π)) = dxj
so
(ia ◦ π)∗ ω = ωJ (x, a) dxJ .
Thus,
d ◦ P (ω) + P ◦ d (ω) = 0 + ω − (ia ◦ π)∗ ω
as required.
For the second case one has, as above, (ia ◦ π)∗ (ωI (x, t)) = ωI (x, a) and
(ia ◦ π)∗ (dxi ) = dxi , but now
so
(ia ◦ π)∗ ω = 0.
Next, [∫ ]
t
Pω = ωI (x, s) ds dxI
a
so
[( ∫ t ) ( ∫ t ) ]
∂ ∂
d ◦ P (ω) = ωI (x, s) ds dx + α
ωI (x, s) ds dt ∧ dxI
∂xα a a ∂t
[∫ t ]
∂ωI
= (x, s) ds dxα ∧ dxI + ωI (x, t) dt ∧ dxI
a ∂xα
[∫ t ]
∂ωI
= (x, s) ds dxα ∧ dxI + ω.
a ∂xα
270 5. de Rham Cohomology
and
π # : Hdek R (X) −→ Hdek R (X × R)
are inverses of each other for every k = 0, 1, 2, . . .. In particular,
Hdek R (X × R) ∼
= Hdek R (X).
so ω − (ia ◦ π)∗ ω is exact. Thus, [ω] = [(ia ◦ π)∗ ω] = (ia ◦ π)# ([ω]) so
(ia ◦ π)# = idHdek R (X×R) , i.e., π # ◦ ia# = idHdek R (X×R) and the result follows.
Remark: Note that it follows from Corollary 5.2.2 that, if a and b are any
two real numbers, then
ia# = ib# .
h# ◦ (h′ )# = idHde
k
R (X)
for any k so h# and (h′ )# are inverse isomorphisms.
In particular, we have the following very useful computational device.
Corollary 5.2.4 Two manifolds of the same smooth homotopy type have the
same de Rham cohomology groups.
Remark: From the Remarks following Corollary 3.2.3 it follows that the
de Rham cohomology groups are homotopy invariants (and, in particular,
homeomorphism invariants).
A manifold X is smoothly contractible if the identity map idX is
smoothly homotopic to a constant map from X to X.
Exercise 5.2.4 Show that X is smoothly contractible if and only if it has the
same smooth homotopy type as a point (connected, 0-dimensional manifold).
Hint: The corresponding topological result is Theorem 2.3.7 of [N4].
Corollary 5.2.5 A smoothly contractible manifold X has trivial de Rham
R (X) for all k ≥ 1.
k
cohomology groups Hde
Let X be a manifold and A a smooth submanifold of X. Then the
inclusion map ι : A ,→ X is smooth (in fact, an embedding, by Lemma 5.6.1
of [N4]). A smooth map r : X → A is called a smooth retraction of X
onto A if r ◦ ι = idA . In this case, ι# ◦ r# = idHde
k
R (A)
for each k. A smooth
retraction r is called a smooth deformation retraction if ι ◦ r : X → A
is smoothly homotopic to idX . In this case, A and X obviously have the
same smooth homotopy type and therefore isomorphic cohomology groups,
but more is true.
Exercise 5.2.5 Show that, if r : X → A is a smooth deformation retraction
of X onto A, then ι# and r# are inverse isomorphisms for each k.
Exercise 5.2.6 Show that there is a smooth deformation retraction of
R2 − { (0, 0) } onto S 1 and then describe all of the de Rham cohomology
groups of R2 − { (0, 0) }.
Corollary 5.2.2 generalizes easily to the case in which R is replaced by
any smoothly contractible manifold. Specifically, we suppose X and Y are
manifolds with Y smoothly contractible. Let F : Y × R → Y be a smooth
map with F (y, 0) = y and F (y, 1) = y0 ∈ Y for every y ∈ Y . We also let
π : X × Y → X be the projection and iy0 (x) = (x, y0 ). Then π ◦ iy0 = idX .
Exercise 5.2.7 Show that iy0 ◦ π is smoothly homotopic to idX×Y .
Hint: Define H : (X × Y ) × R → X × Y by H((x, y), t) = (x, F (y, t)).
It follows that π # : Hde R (X) → Hde R (X × Y ) and iy0 R (X × Y ) →
k k # k
: Hde
k
Hde R (X) are inverse isomorphisms and, in particular, we have the following
result.
Theorem 5.2.6 Let X and Y be smooth manifolds with Y smoothly con-
∼ k
R (X × Y ) = Hde R (X) for every k.
k
tractible. Then Hde
272 5. de Rham Cohomology
One more result along these lines will be of use. We wish to consider a
smooth vector bundle over X. More precisely, we begin with a
P
smooth principal G-bundle G ,→ P → X over X and a representation
ρ : G → GL(V) of G on some finite dimensional vector space V. Now consider
the associated vector bundle
Pρ : P ×ρ V −→ X.
We will show that P ×ρ V and X have the same de Rham cohomology groups.
For this recall that each fiber Pρ−1 (x) = { [p, v] : v ∈ V }, where p is any point
in P −1 (x), has a natural vector space structure under which it is isomorphic
to V. In particular, each Pρ −1 (x), has a copy [p, 0] of the zero element in V
and we may define a map σ0 : X → P ×ρ V by σ0 (x) = [p, 0].
Exercise 5.2.8 Show that σ0 is a smooth global cross-section of P ×ρ V.
The cross-section σ0 is called the zero section of P ×ρ V and, because it is a
cross-section, it satisfies Pρ ◦ σ0 = idX . Defining F : (P ×ρ V) × R → P ×ρ V
by F ([p, v], t) = [p, tv] we obtain a smooth map with F ([p, v], 0) = [p, 0] = σ0 ◦
Pρ ([p, v]) and F ([p, v], 1) = [p, v] = idP ×ρ V ([p, v]). Thus, σ0 ◦ Pρ is homotopic
to idP ×ρ V . We conclude that σ0 # : Hde R (P ×ρ V) → Hde R (X) and Pρ
k k #
:
Hde R (X) → Hde R (P ×ρ V) are inverse isomorphisms.
k k
Theorem 5.2.7 A smooth vector bundle and its base manifold have the same
de Rham cohomology groups.
The situation is very different for principal bundles, however, as we shall see
in Section 5.4.
Corollary 5.2.4 is an important tool in the calculation of cohomology groups,
but its effective use requires a means of putting together the cohomology of
a manifold X from that of two open submanifolds of which it is the union.
For example, S 2 = UN ∪ US , where UN and US are S 2 minus the south and
north poles, respectively. Both UN and US are diffeomorphic to R2 and so
have trivial cohomology. Moreover, UN ∩US is diffeomorphic to the punctured
plane R2 −{ (0, 0) } and so we know its cohomology as well (because it has the
same homotopy type as S 1 ). What’s needed then is some relationship between
R (UN ∪US ), Hde R (UN ), Hde R (US ) and Hde R (UN ∩US ). This is provided
k k k k
Hde
by the “Mayer-Vietoris sequence” which is our next major objective. We take
one preliminary step to conclude this section, an important algebraic detour
in the next and finally prove the theorem in Section 5.4. This will all seem
quite familiar to those who have been through the proof of Mayer-Vietoris
for singular homology in Section 3.5 of [N4], but the argument here is much
simpler and one should keep an eye out to spot the precise point at which one
can account for this relative simplicity.
We consider then a smooth manifold X = U ∪ V , where U and V are open
submanifolds of X. Introduce the inclusions
5.2. Induced Homomorphisms 273
iU : U ,→ X
iV : V ,→ X
jU : U ∩ V ,→ U
jV : U ∩ V ,→ V
We also define
αk = iU ∗ ⊕ iV ∗ : Λk (X) −→ Λk (U ) ⊕ Λk (V )
by
αk (ω) = (iU ∗ ω, iV ∗ ω)
and
β k = jU ∗ − jV ∗ : Λk (U ) ⊕ Λk (V ) −→ Λk (U ∩ V )
by
β k (λ1 , λ2 ) = jU ∗ λ1 − jV ∗ λ2 .
Thus, for each k, we have a sequence
αk βk
0 −→ Λk (X) −→ Λk (U ) ⊕ Λk (V ) −→ Λk (U ∩ V ) −→ 0 (5.2.9)
which we claim is exact (i.e., the image of each map is the kernel of the next).
Exactness at Λk (X) is the statement that αk is one-to-one. To see this suppose
ω 1 and ω 2 are distinct elements of Λk (X). Since X = U ∪ V , there is a p in
U or in V (or in both) at which ω 1 (p) ̸= ω 2 (p). Since U and V are open in
X, the tangent space at p to either U or V coincides with Tp (X) so either
iU ∗ ω 1 (p) ̸= iU ∗ ω 2 (p) if p ∈ U or iV ∗ ω 1 (p) ̸= iV ∗ ω 2 (p) if p ∈ V . In either
case, αk (ω 1 ) ̸= αk (ω 2 ).
To prove exactness at Λk (U ) ⊕ Λk (V ) we must show that Image (αk ) =
ker (β k ) and for this we prove containment in each direction. Since αk (ω) =
(iU ∗ ω, iV ∗ ω), β k ◦ αk (ω) = jU ∗ (iU ∗ ω) − jV ∗ (iV ∗ ω) = 0 because both terms
are the restrictions of ω to U ∩ V . Thus, Image (αk ) ⊆ ker (β k ). Next we show
that ker (β k ) ⊆ Image (αk ). Suppose β k (λ1 , λ2 ) = 0. Then jU ∗ (λ1 ) = jV ∗ (λ2 )
so λ1 and λ2 agree on U ∩ V . Thus, we can define ω on X = U ∪ V by taking
iU ∗ ω = λ1 and iV ∗ ω = λ2 . Then αk (ω) = (λ1 , λ2 ) so (λ1 , λ2 ) ∈ Image (αk ).
274 5. de Rham Cohomology
β k (λ1 , λ2 ) = jU ∗ λ1 − jV ∗ λ2 = ϕV ω − (−ϕU ω)
= (ϕV + ϕU ) ω = ω.
so ω ∈ Image (β k ) as required.
Thus, we have an exact sequence (5.2.9) for each k. Letting d denote
the exterior differentiation operator on all of the spaces of forms (Λk (X),
Λk (U ), Λk (V ) or Λk (U ∩ V ) for any k) we define
d ⊕ d : Λk (U ) ⊕ Λk (V ) −→ Λk+1 (U ) ⊕ Λk+1 (V )
d d%d d
αk βk
0 Λk (X) Λk (U) % Λk (V) Λk (U ∩ V) 0
d d%d d
αk + 1 βk + 1
0 Λk + 1 (X) Λk + 1 (U) % Λk + 1 (V) Λk + 1 (U ∩ V) 0
d d%d d
We have just shown that the diagram has exact rows. The columns, although
not exact, in general, compose to zero at each stage because d2 = 0.
Exercise 5.2.9 Show that each square in this diagram is commutative, i.e.,
that, for each k,
δ k−1 δk δ k+1
· · · −→ C k−1 −→ C k −→ C k+1 −→ · · ·
defined for all integers k such that the image of each homomorphism is
contained in the kernel of the next, i.e., δ k+1 ◦δ k = 0 for each k. In the only ex-
ample we have seen thus far C k was the R-module Λk (X) of k-forms on a man-
ifold X and δ k was the exterior differentiation map dk . The homomorphism
δ k : C k → C k+1 is called the k th coboundary operator and, when it is con-
venient and no confusion will result, will be denoted simply δ. Image (δ k−1 ) is
a submodule of C k and its elements are called k -coboundaries. ker (δ k ) is a
submodule of C k and its elements are called k -cocycles. Since δ k ◦ δ k−1 = 0,
276 5. de Rham Cohomology
α : C1∗ −→ C2∗
such that
δ2k ◦ αk = αk+1 ◦ δ1k
for each k.
αk
C1k C2k
δ1k δ2k
αk + 1
C1k + 1 C2k + 1
Exercise 5.3.1 Show that αk (ker (δ1k )) ⊆ ker (δ2k ) and αk (Image (δ1k−1 )) ⊆
Image (δ2k−1 ) for each k and conclude that αk induces a homomorphism
in cohomology defined by
in de Rham cohomology.
Exercise 5.3.2 Two cochain maps α : C1∗ → C2∗ and β : C1∗ → C2∗ are said
to be algebraically homotopic (or cochain homotopic) if there exists a
family of homomorphisms
such that
hk+1 ◦ δ2k + δ1k−1 ◦ hk = αk − β k
(αk )# = (β k )#
for each k. Hint: The argument is the same as for de Rham cohomology.
The composition of two cochain maps C1∗ → C2∗ and C2∗ → C3∗ is defined by
composing each of the homomorphisms C1k → C2k and C2k → C3k and clearly
induces homomorphisms in cohomology that are just the compositions of those
induced by the two cochain maps. Denoting by 0 both the trivial R-module
and the cochain complex one can form from these we will say that a sequence
of cochain maps
β
0 −→ C1∗ −→ C2∗ −→ C3∗ −→ 0
α
forms a short exact sequence if, for each k, the sequence of R-modules
αk βk
0 −→ C1k −→ C2k −→ C3k −→ 0
is exact (i.e., the image of each map equals the kernel of the next).
278 5. de Rham Cohomology
αk βk
0 C1k C2k C3k 0
αk + 1 βk + 1
0 C1k + 1 C2k + 1 C3k + 1 0
Now, α and β both induce maps in cohomology, but it need not be the case
that the sequences
∂ k−1 (αk )# (β k )# ∂k
· · · H k−1 (C3∗ ) −→ H k (C1∗ ) −→ H k (C2∗ ) −→ H k (C3∗ ) −→ H k+1 (C1∗ ) −→ · · ·
Proof: To define ∂ k : H k (C3∗ ) → H k+1 (C1∗ ) we will make two choices and
then prove that the definition is independent of those choices and has the
required properties.
Any element of H k (C3∗ ) is [x] for some x ∈ C3k with δ3k (x) = 0. Since β k is
surjective, there exists a y ∈ C2k with β k (y) = x. Now, observe that
( ) ( )
β k+1 δ2k (y) = δ3k β k (y) = δ3k (x) = 0
5.3. Cochain Complexes and Their Cohomology 279
so
δ2k (y) ∈ ker (β k+1 ) = Image (αk+1 ).
Since αk+1 is injective, there is a unique z ∈ C1k+1 such that
and so
δ2k (y − αk (u)) = 0.
Let y ′ = y − αk (u) ∈ C2k . Then δ2k (y ′ ) = 0 and β k (y ′ ) = β k (y − αk (u)) =
β k (y) − β k (αk (U )) = β k (y) = x as required.
are defined by
( ) ( )
δ k (x1 , x2 ) = δ1k ⊕ δ2k (x1 , x2 ) = δ1k (x1 ), δ2k (x2 ) .
Exercise 5.3.5 Show that C1∗ ⊕ C2∗ is, indeed, a cochain complex and that,
for each k,
H k (C1∗ ⊕ C2∗ ) ∼
= H k (C1∗ ) ⊕ H k (C2∗ ).
5.4. The Mayer-Vietoris Sequence 281
such that the following long sequence (eventually ending with zeros) is exact:
(α0 )# (β 0 )#
0 −→ Hde
0
R (X) −→ Hde R (U ) ⊕ Hde R (V ) −→ Hde R (U ∩ V )
0 0 0
∂0 ∂ k−1 (αk )#
−→ · · · Hde
k−1
R (U ∩ V ) −→ Hde R (X) −→ Hde R (U ) ⊕ Hde R (V )
k k k
(β k )# ∂k
−→ Hde
k
R (U ∩ V ) −→ Hde R (X) −→ · · · .
k+1
0 −→ Hde
k
R (X) −→ Hde R (U ) ⊕ Hde R (V ) −→ 0
k k
Proof: The proof is by induction on the size of the simple cover. If X has
a simple cover consisting of just one open set, then it is diffeomorphic to
Rn (n = dim X) so the result follows from Theorem 5.1.1. Assume now that
the result has been established for manifolds admitting simple covers consist-
ing of N open sets and suppose that X has a simple cover { U1 , . . . , UN , UN +1 }
consisting of N +1 open sets. Let U = U1 ∪· · ·∪UN and V = UN +1 . Then U ∩V
is an open submanifold of X and { U1 ∩ UN +1 , . . . , UN ∩ UN +1 } is a simple
cover of it. By the induction hypothesis, all of the cohomology groups of U and
U ∩ V are finite dimensional. Obviously, the same is true of V = UN +1 ∼ = Rn .
Now, X = U ∪ V so, for each k, the Mayer-Vietoris sequence gives the exact
sequence
∂ α
· · · −→ Hde
k−1
R (U ∩ V ) −→ Hde R (X) −→ Hde R (U ) ⊕ Hde R (V ) −→ · · ·
k k k
0 −→ Hde
0
R (S ) −→ Hde R (UN ) ⊕ Hde R (US ) −→ Hde R (UN ∩ US ) −→
1 0 0 0
i.e.,
0 −→ R −→ R ⊕ R −→ R ⊕ R −→ H 1 (S 1 ) −→ 0 −→ 0 −→ 0
α β ∂
0 −→ R −→ R ⊕ R −→ R ⊕ R −→ H 1 (S 1 ) −→ 0.
0 −→ V1 −→ V2 −→ V3 −→ · · · −→ Vk−1 −→ Vk −→ 0
α β
0 −→ V1 −→ V2 −→ V3 −→ · · · −→ Vk−1 −→ Vk −→ 0.
β̂ : V2 /Image (α) −→ V3
βˆ
0 −→ V2 /Image (α) −→ V3 −→ · · · −→ Vk −1 −→ Vk −→ 0
x − xn+1 en+1
r(x) =
|x − xn+1 en+1 |
N
x ∈ UN ∩ US
xn + 1 en + 1
en + 1
r(x)
Sn – 1
x – xn + 1 en + 1
S
5.4. The Mayer-Vietoris Sequence 285
txn + 1 en + 1 H (x, t)
r (x)
x – txn + 1 en + 1
Now, to carry out an induction one needs an induction hypothesis. Our only
information thus far concerns S 1 :
{
k 1 ∼ R, if k = 0, 1
Hde R (S ) = .
0, if k ̸= 0, k ̸= 1
To check that this is indicative of the general result (and to get a bit more
2 ∼
practice with Mayer-Vietoris) we will do S 2 as well. Of course, Hde
k
R (S ) = 0
for k ≥ 3 so the Mayer-Vietoris sequence gives
0 −→ Hde
0
R (S ) −→ Hde R (UN ) ⊕ Hde R (US ) −→ Hde R (UN ∩ US ) −→
2 0 0 0
0 −→ R −→ R ⊕ R −→ R −→ Hde
1
R (S ) −→ 0
2
and
0 −→ R −→ Hde
2
R (S ) −→ 0.
2
286 5. de Rham Cohomology
1 − 2 + 1 − dim (Hde
1 2
R (S )) = 0
and
1 − dim (Hde
2 2
R (S )) = 0
so
2 ∼ 2 ∼
R (S ) = 0 and Hde R (S ) = R.
1 2
Hde
We conclude then that
{
k 2 ∼ R, if k = 0, 2
Hde R (S ) = .
0, if k ̸= 0, k ̸= 2
Remark: Notice that, for k = 1, the sequence of forms and the corres-
ponding cohomology sequence are
0 −→ 0 −→ 0 ⊕ 0 −→ R −→ 0
which certainly is not. This provides the example promised just before
Theorem 5.3.1.
Exercise 5.4.3 Prove, by induction on n ≥ 1, that
{
k n ∼ R, if k = 0, n
Hde R (S ) = .
0, if k ̸= 0, k ̸= n
so ∫ ∫
ν =α ω.
Sn Sn
∫ ∫
If S n ν = 0, then, since S n ω ̸= 0, we must have α = 0 and so [ν] = [0], i.e.,
ν is exact.
Remark: We will eventually show that the same is true for any compact,
orientable manifold.
Exercise 5.4.4 For [ω] ∈ Hde
n n n
R (S ), define the integral of [ω] over S by
∫ ∫
[ω] = ω.
Sn Sn
Exercise 5.4.5 Let 0 denote the origin (0, . . . , 0) in Rn+1 , n ≥ 1. Show that
{
R, if k = 0, n
k
Hde R (R
n+1
− {0}) ∼
= .
0, if k ̸= 0, k ̸= n
have derivatives that are isomorphisms. The game here is to get from the
smooth to the topological category. This was done using singular homology in
Section 3.4 of [N4], but the computations required to do it were substantially
more involved.
by
h# = f # ,
where f is any smooth map homotopic to h. We ask the reader to establish
all of the usual properties.
Exercise 5.4.7 Let U, V and W be open sets in Euclidean spaces and
k any integer. Prove each of the following.
1. If h0 , h1 : U −→ V are homotopic continuous maps, then
h# #
0 = h1 : Hde R (V ) −→ Hde R (U ).
k k
U V
U∩V
Note that ∫ ∫ ∫
−1 ∗
(ϕ ) ω= ω= ω =1
Rn U X
so ∫
(ϕ−1 )∗ (c ω ) = c.
Rn
292 5. de Rham Cohomology
Thus, ∫
(ϕ−1 )∗ (ξ − c ω ) = 0.
Rn
Since (ϕ−1 )∗ (ξ − c ω ) has compact support, Lemma 5.5.1 implies that there
exists a compactly supported η ∈ Λn−1 (Rn ) such that
(ϕ−1 )∗ (ξ − c ω ) = dη
on Rn . Thus,
ξ − c ω = ϕ∗ (dη) = d(ϕ∗ η)
on U . Now, ϕ∗ η has compact support contained in U so, with a bump function
that is 1 on supp (ϕ∗ η) and 0 outside U , we may regard it as an (n − 1)-form
ν on all of X. Then
ξ − c ω = dν
on all of X (because everything is zero outside of U ) so ξ and c ω are coho-
mologous on X.
Theorem 5.5.4 Let X be a compact. connected, orientable n-manifold. Then
∼
R (X) = R.
n
Hde
U = V1 , V2 , . . . , Vk−1 , Vk = V
is an isomorphism.
Exercise 5.5.5 Prove Corollary 5.5.6.
Exercise 5.5.6 Suppose X is a compact, connected orientable n-manifold,
where n = 2k for some k ≥ 1. Define QX : Hde
k
R (X) × Hde R (X) −→ R by
k
∫
QX ([ξ], [η]) = ξ ∧ η.
X
bk = dim Hde
k
R (X)
Although it is not obvious, this alternating sum coincides with the Euler
characteristic of X, defined in Chapter 3 of [N4] as the alternating sum of
the ranks of the singular homology groups of X. We will have a bit more to
say about the Euler characteristic in the Appendix.
We must show that this definition does not depend on the choice of the nor-
n ′
malized ∫generator ω 0 for Hde R (Y ). Thus, suppose ω 0 is another n-form on
′
Y with Y ω 0 = 1. Since [ω 0 ] generates Hde R (Y ) there exists an α ∈ R such
n
that
ω ′0 = αω 0 + dη
ω = αω 0 + dη
which we write as ∫ ∫
∗
f ω = deg(f ) ω. (5.6.2)
X Y
Concrete calculations are greatly facilitated by the fact that deg is a homotopy
invariant.
Theorem 5.6.1 Let X and Y be compact, connected orientable n-manifolds
and suppose f, g : X −→ Y are two smooth maps that are smoothly homotopic.
Then
deg(f ) = deg(g).
n
Thus, given any normalized generator ω 0 for Hde R (Y ) we have
# # ∗ ∗
f ([ω 0 ]) = g ([ω 0 ]) so [f ω 0 ] = [g ω 0 ] and, consequently,
f ∗ ω 0 = g ∗ ω 0 + dη
f : X −→ S n
W ′ = f −1 (W ) − (U1 ∪ · · · ∪ Uk ).
This is closed in X and therefore compact. Consequently, f (W ′ ) is a closed
set in Y that does not contain q. Select a coordinate neighborhood V of q with
V ⊆ W − f (W ′ ). Then f −1 (V ) ⊆ U1 ∪ · · · ∪ Uk . Now, if necessary, replace
each Ui by Ui ∩ f −1 (V ) (still a coordinate neighborhood) so that
f −1 (V ) = U1 ∪ · · · ∪ Uk .
Since V must contain an open set diffeomorphic to R∫n , Lemma 5.5.2 gives
an n-form ω 0 on Y with support contained in V and Y ω 0 = 1. Moreover,
supp (f ∗ ω 0 ) ⊆ U1 ∪ · · · ∪ Uk so, since this union is disjoint,
∫ k ∫
∑
f ∗ ω0 = f ∗ ω0 .
X i=1 Ui
where the plus sign is chosen if f is orientation preserving on Ui and the minus
sign is chosen if f is orientation reversing on Ui . Thus,
∫ {
∗ 1 if f∗pi is orientation preserving
f ω0 =
Ui −1 if f∗pi is orientation reversing
= sign(f, pi )
To see that the definition does not depend on the choice of ω, suppose ω ′ is
another (n − 1)-form on S 2n−1 with dω ′ = f ∗ ω 0 . Then d(ω − ω ′ ) = 0 so
∫ ∫ ∫
′ ′
ω ∧ dω − ω ∧ dω = (ω − ω ′ ) ∧ dω
S 2n−1 S 2n−1
∫ S 2n−1
= d ((ω − ω ′ ) ∧ ω ) = 0
S 2n−1
by Stokes’ Theorem.
Exercise 5.7.1 Show that if n is odd, then H(f ) = 0 for any smooth map
f : S 2n−1 −→ S n . Hint: Compute d(ω ∧ ω) for any even dimensional form ω.
We show now that the Hopf invariant is a smooth homotopy invariant.
Theorem 5.7.1 Let f0 , f1 : S 2n−1 −→ S n , n ≥ 2, be smoothly homotopic
maps. Then H(f0 ) = H(f1 ).
300 5. de Rham Cohomology
and
ι1 : S1 = S 2n−1 × {1} ,→ S 2n−1 × (0 − ϵ, 1 + ϵ)
be the inclusion maps. Identifying S0 and S1 with S 2n−1 we may also identify
F ◦ ι0 with f0 and F ◦ ι1 with f1 , respectively. Now let ω 0 be a normalized
∗
generator for Hde n n
R (S ). Then F ω 0 is an n-form on S
2n−1
× (0 − ϵ, 1 + ϵ). By
n
Theorem 5.2.6, Hde R (S 2n−1
× (0 − ϵ, 1 + ϵ)) ∼ n
= Hde R (S 2n−1
) and, since n ≥ 2,
this is trivial. Consequently, there is an (n − 1)-form η on S 2n−1 × (0 − ϵ, 1 + ϵ)
such that
F ∗ ω 0 = dη.
Let ι∗0 η = ω and ι∗1 η = ω ′ . Notice that
and, similarly,
f1∗ ω 0 = dω ′ .
Moreover,
and, similarly,
ω ′ ∧ dω ′ = ι∗1 (η ∧ dη).
We must show that ∫ ∫
ω ∧ dω = ω ′ ∧ dω ′ ,
So S1
i.e., that ∫ ∫
ι∗0 (η ∧ dη) = ι∗1 (η ∧ dη). (5.7.4)
S0 S1
d(η ∧ dη) = dη ∧ dη = F ∗ ω 0 ∧ F ∗ ω 0 = F ∗ (ω 0 ∧ ω 0 ) = 0
Remark: As was the case for the Brouwer degree there is another, more
geometrical way of calculating the Hopf invariant which shows that H(f ) is
actually an integer. This involves the linking number of the preimages f −1 (p)
and f −1 (q) of any two distinct regular values of f (see Chapter III, Section 17,
of [BT]). Still other approaches to the Hopf invariant are described in [Huse].
We conclude with a few exercises that will lead the reader through the
calculation of the Hopf invariant for the projection map
P1 : S 3 −→ S 2
of the complex Hopf bundle. The result will be H(P1 ) = 1 so that, in partic-
ular, Exercise 5.7.3 then implies that P1 is not nullhomotopic. This was, in
fact, the first example of a homotopically nontrivial map of one sphere onto
another of smaller dimension and came as quite a shock in the 1930s when
Hopf constructed it. Recall that if S 3 is identified with the subset of C2 con-
sisting of those (z1 , z2 ) for which |z1 |2 + |z2 |2 = 1 and S 2 is identified with a
subset of R3 , then
( )
P1 (z1 , z2 ) = z 1 z̄ 2 + z̄ 1 z 2 , −i z 1 z̄ 2 + iz̄ 1 z 2 , |z 1 |2 − |z 2 |2 .
For the calculation of the Hopf invariant we will take the normalized gener-
n 2 2
ator ω 0 of Hde R (S ) to be the standard volume form for S divided by the
“volume” 4π of S (see Section 4.6). Recall that, if x , x and x3 are standard
2 1 2
6.1 Motivation
We have had a number of previous, albeit informal encounters with the notion
of a characteristic class for a principal bundle (in Section 2.2, 2.4 and 2.5) and
are now in possession of sufficient machinery to take up the subject in earnest.
Before plunging into the thick of the battle, however, we would like to make
clear the strategy we propose to adopt. This is perhaps best achieved by
going through what might be regarded as the “trivial” case in detail and then
indicating the modifications required when matters are not so simple.
P
We begin then by considering a principal U (1)-bundle U (1) ,→ P −→ X
over a manifold X and let ω denote a connection on the bundle. Thus,
one might have in mind an electromagnetic field on spacetime. Since U (1)
is Abelian, the curvature is given by Ω = dω. For any local cross-section
s : V −→ P −1 (V ) we write, as usual, A = s∗ ω = −i A and F =
s∗ Ω = −i F for the local gauge potential and field strength. Again be-
cause U (1) is Abelian, the field strengths F for various cross-sections s
agree on the intersections of their domains and thereby determine a glob-
ally defined u(1)-valued 2-form on X that we also denote by F .
This F is −i F for a globally defined real-valued 2-form F on X. Since
1
dF = 0, F is closed and therefore 2π F determines an element of
Hde2
R (X; R).
Now suppose ω ′ is another connection on the same bundle
P
U (1) ,→ P −→ X determining global 2-forms F ′ and F ′ in the same way. It
follows that ω − ω ′ = τ is in Λ1ad (P, u(1)). Thus, dω − dω ′ = dτ so
Ω − Ω ′ = dτ .
Since τ is tensorial of type ad and U (1) is Abelian, σg∗ τ = g −1 ·τ = gτ g −1 = τ
for every g ∈ U (1). But then Lemma 4.5.1. implies that τ projects to X, i.e.,
that there exists a unique u(1)-valued 1-form τ̄ on X such that τ = P ∗ τ̄ .
Since (in the Abelian case) curvatures also project to X, Ω = P ∗ F and
Ω ′ = P ∗ F ′ and therefore P ∗ F − P ∗ F ′ = d(P ∗ τ̄ ) so
P ∗ (F − F ′ ) = P ∗ (dτ̄ ).
But projections to X are unique when they exist so we must have
F − F ′ = dτ̄ .
1
In particular, 2π 1
F and 2π F ′ are cohomologous and therefore determine the
same element of Hde R (X; R). The unique cohomology class determined in
2
P
this way from any connection on U (1) ,→ P −→ X is called the 1 st Chern
class of the bundle and is denoted
[ ]
1
F ∈ Hde R (X; R).
2
c1 (P ) =
2π
only that we have some connection on the bundle with local gauge potentials
AN and AS on UN and US , respectively. Then, on UN ∩ US ,
−1 −1
AN = gSN AS gSN + gSN dgSN
so
AN = AS − i n d θ.
∫ ∫
i i
= ι∗ AN − ι∗ AS
2π S 1 2π S 1
∫ ∫
i i
= ι∗ (AS − i n d θ) − ι∗ AS
2π S 1 2π S 1
∫ ∫ ∫
i n i
= ι∗ AS + ι∗ d θ − ι∗ AS
2π S 1 2π S 1 2π S 1
∫
n
= d(θ ◦ ι).
2π S 1
The essential feature of the calculation to this point is that all references to
the connection itself have dropped out and we are left with an integral that in-
−1
volves only the transition function (gSN dgSN = −i n d θ). This is as it should
be, of course, since the Chern class (and therefore its integral) depends only
on the bundle and not on the connection. Calculating the remaining integral
is easy and, in fact, was done explicitly in Section 4.6. The result was simply
the length (“volume”) of S 1 :
∫
d(θ ◦ ι) = 2π.
S1
Thus ∫
n
c1 (P ) = (2π) = n
S2 2π
as promised.
This then concludes our discussion of the “trivial” case and the issue before
us now is whether or not it is possible to push through an analogous program
306 6. Characteristic Classes
for bundles with other structure groups. On the surface, the prospects do not
appear to be good since virtually every stage in our construction depended on
the commutativity of U (1). It is not even altogether clear how to get the ball
rolling since the very existence of the 2-form on X whose cohomology class
contained the topological information about the bundle required that local
field strengths agree on the intersections of their domains. Since, in general,
such local field strengths are related by F g = g −1 F g, this simply is not
true when the structure group is non-Abelian. The key idea here is to evade
this difficulty by considering, not the field strengths themselves, but various
functions of the field strength that cannot tell the difference between F and
g −1 F g. Identifying all of these objects with matrices there are a number of
obvious choices. For example, since the trace of a matrix is invariant under
conjugation, trace (g −1 F g) = trace F and the 2-forms defined locally on X
by taking the trace of the local field strengths will agree on the intersections
of their domains and therefore determine a globally defined 2-form trace F on
X. Perhaps trace F is closed and so determines a cohomology class. Perhaps
this cohomology class is independent of the connection from which trace F
was constructed. Perhaps the same is true for other ad-invariant functions of
F such as trace (F ∧ F ), or det F . Perhaps this is all a bit too much to ask.
We shall see.
Exercise 6.2.1 Show that f˜ ⊙ g̃ ∈ S k+l (V). Extend ⊙ to S(V) and show
that, with this operation, S(V) has the structure of a commutative algebra
with identity.
If {e1 , . . . , en } is a basis for V and {e1 , . . . , en } is the dual basis for V ∗ , then
for any f˜ ∈ S k (V), there exist unique complex numbers ai1 ...ik , i1 , . . . , ik =
1, . . . , n, symmetric in i1 , . . . , ik , such that
Exercise 6.2.3 Show that, if f ∈ P (V) and g ∈ P l (V), then the product f g
k
1[ ]
f˜(v, w) = f (v + w) − f (v) − f (w)
2
for all v, w ∈ V. Indeed, one need only compute f˜(v+w, v+w) using bilinearity
to obtain
f (v + w) = f (v) + 2f˜(v, w) + f (w).
1[
f˜(u, v, w) = f (u + v + w) − f (u + v) − f (u + w)
6 ]
− f (v + w) + f (u) + f (v) + f (w) .
Thus, f˜(v, w) is 1
2 times the coefficient of st in the expansion of f (sv + tw).
Exercise 6.2.4 Show that, in the general case, f˜(v1 , . . . , vk ) is k! 1
times the
coefficient of t1 t2 · · · tk in the expansion of f (t1 v1 + t2 v2 + · · · + tk vk ).
The linear isomorphisms f˜ −→ f of S k (V) onto P k (V), together with
the identity map of S 0 (V) onto P 0 (V), gives a linear isomorphism of S(V)
onto P (V). To show that this is, in fact, an algebra isomorphism it will suf-
fice to prove that, if f˜ ∈ S k (V) with f˜ −→ f ∈ P k (V) and g̃ ∈ S l (V)
6.2. Algebraic Preliminaries 309
as required.
Now, suppose that G is a Lie group and ρ : G −→ GL(V) is a representation
of G on V. For each k ≥ 1, let Sρk (V) and Pρk (V) denote the subspaces of S k (V)
and P k (V), respectively, that are invariant under ρ, i.e., satisfy
and
f (ρ(g)(v)) = f (v)
defined by
1 ∑ ( )
symtr(A1 , . . . , Ak ) = trace Aσ(1) · · · Aσ(k) . (6.2.2)
k!
σ∈Sk
310 6. Characteristic Classes
(the reason for the i will emerge shortly; the reason for the 2 π, a bit later).
When expanded this determinant is a polynomial in λ of degree n if the
dimension of G is n. The coefficients of this polynomial depend on A and we
write them as follows:
( )
i
det λI + A = f0 (A)λn + f1 (A)λn−1 + · · · + fn−1 (A)λ + fn (A). (6.2.4)
2π
The functions fk (A) are invariant under conjugation since
( ) ( ( ) )
i −1 i
det λI + (g Ag) = det g −1 (λI )g + g −1 A g
2π 2π
( ( ) )
i
= det g −1 λI + A g
2π
( )
i
= det λI + A .
2π
Notice that f0 (A) = 1 for any A. To obtain convenient expressions for the
remaining fk (A) and show that they are, in fact, homogeneous polynomials
we recall a few basic facts concerning the elementary symmetric functions.
Let F denote a field of characteristic zero (e.g., R or C) and denote by
F[x1 , . . . , xn ] the algebra of polynomials in the n variables x1 , . . . , xn with
coefficients in F. For any P ∈ F[x1 , . . . , xn ] and σ ∈ Sn define σ P by
σ
P (x1 , . . . , xn ) = P (xσ(1) , . . . , xσ(n) ). Then P is symmetric if σ P = P for
every σ ∈ Sn . The elementary symmetric polynomials S0 , S1 , . . . , Sn are
the elements of F[x1 , . . . , xn ] defined by
6.2. Algebraic Preliminaries 311
S0 (x1 , x2 , . . . , xn ) = 1
S1 (x1 , x2 , . . . , xn ) = x1 + x2 + · · · + xn
S2 (x1 , x2 , . . . , xn ) = x1 x2 + x1 x3 + · · · + x1 xn +
x2 x3 + · · · + x2 xn +
· · · + xn−1 xn
S3 (x1 , x2 , . . . , xn ) = x1 x2 x3 + x1 x2 x4 + · · · + x1 x2 xn +
x1 x3 x4 + · · · + x1 x3 xn + · · · + xn−2 xn−1 xn
..
.
Sk (x1 , x2 , . . . , xn ) = x1 x2 · · · xk + · · · + xn−k+1 xn−k+2 · · · xn
..
.
Sn (x1 , x2 , . . . , xn ) = x1 x2 · · · xn .
f0 (A) = 1. (6.2.7)
S2 (a1 , . . . , an ) = a1 a2 + · · · + a1 an + a2 a3 + · · · + a2 an + · · · + an−1 an
1[ ]
= (a1 + · · · + an )2 − (a21 + · · · + a2n ) .
2
Again, a1 + · · · + an = trace A. Furthermore,
(( )2 ) ( )
a21 + · · · + a2n = trace g −1 Ag = trace (g −1 Ag)(g −1 Ag)
= trace(g −1 (A2 )g) = trace (A2 )
so we have
1[ ]
S2 (a1 , . . . , an ) = (trace A)2 − trace(A2 )
2
and therefore
1 [ ]
f2 (A) = − (trace A) 2
− trace(A 2
) . (6.2.9)
8π 2
1[ ]
S3 (a1 , . . . , an ) = (trace A)3 − 3 trace(A2 )trace A + 2 trace(A3 )
6
and conclude that
i [
f3 (A) = − (trace A)3
48π 3 ] (6.2.10)
−3 trace(A2 ) trace A + 2 trace(A3 ) .
314 6. Characteristic Classes
There are similar formulas for all of the fk (A), but these will suffice for our
purposes.
At this point we are going to restrict our attention to the two specific Lie
groups that are of most interest to us, i.e., U (n) and SU (n). Observe that, for
either of these, an element A of the Lie algebra is a skew-Hermitian matrix
so i A is Hermitian and therefore has n distinct real eigenvalues. Thus, A has
n distinct pure imaginary eigenvalues a1 , . . . , an . Recalling that each term in
Sk (a1 , . . . , an ) is a product of k distinct elements of {a1 , . . . , an } we have
( )k
i (−1)k
fk (A) = Sk (a1 , . . . , an ) = Sk (i a1 , . . . , i an ).
2π (2π)k
(Chapter XI, [Lang]), there exists a g ∈ U (n) such that g(−i A)g −1 is diago-
nal with real entries, say,
Thus,
−i A = g −1 (diag(ξ 1 , . . . , ξ n ))g
so
A = g −1 diag(i ξ 1 , . . . , i ξ n )g
as required.
Now consider the restriction map R : I(U (n)) −→ I(G′ ), defined as follows:
If f˜: u(n) × · · · × u(n) −→ C is in I(U (n)), then Rf˜ : G ′ × · · · × G ′ −→ C
is given by (Rf˜)(A′1 , . . . , A′k ) = f˜(A′1 , . . . , A′k ) for all A′1 , . . .,A′k in G ′ ⊆ u(n).
Also let
N = {g ∈ U (n) : g −1 A′ g ∈ G ′ for all A′ ∈ G ′ } ⊇ G′
and
{
IN (G′ ) = f˜ ∈ I(G′ ) : f˜(g −1 A′1 g, . . . , g −1 A′k g) =
}
f˜(A′1 , . . . , A′k ) for all A′1 , . . . , A′k in G ′ and all g ∈ N .
We claim that R carries I(U (n)) isomorphically into IN (G′ ). Notice that R
certainly carries I(U (n)) into IN (G′ ) and is an algebra homomorphism. To
show that R is injective we show that its kernel is trivial. Thus, suppose
f˜ ∈ I(U (n)) and Rf˜ is the zero element of IN (G′ ), i.e., (Rf˜)(A′1 , . . . , A′k ) = 0
for all A′1 , . . . , A′k in G ′ . In particular, (Rf˜)(A′ , . . . , A′ ) = 0 for all A′ in G ′ .
By (6.2.12), every A ∈ u(n) can be written as g −1 A′ g for some g ∈ U (n) and
some A′ ∈ G ′ . Thus, f˜(A, . . . , A) = f˜(g −1 A′ g, . . . , g −1 A′ g) = f˜(A′ , . . . , A′ ) =
(Rf˜)(A′ , . . . , A′ ) = 0 for all A ∈ u(n). Consequently, the polynomial on u(n)
corresponding to f˜ is identically zero. Since this correspondence is an isomor-
phism, f˜ is identically zero and the kernel of R is trivial.
To prove that R maps onto IN (G′ ) we note first that every element of
IN (G′ ) gives rise to a polynomial on G ′ that is invariant under conjuga-
tion by elements of N . Identifying G ′ with the set of all diag (i ξ 1 , . . . , i ξ n )
with ξ 1 , . . . , ξ n real, we claim that such a polynomial must be a symmet-
ric function of ξ 1 , . . . , ξ n . To prove this it will suffice to show that, for ev-
ery pair (i, j) of indices with 1 ≤ i < j ≤ n, there exists a g ∈ N such
that
g −1 (diag(i ξ 1 , . . . , i ξ i , . . . , i ξ j , . . . , i ξ n ))g
(6.2.13)
= diag(i ξ 1 , . . . , i ξ j , . . . , i ξ i , . . . , i ξ n ).
Exercise 6.2.7 Show that the n × n matrix g having 1 in the (i, j)-slot, the
(j, i)-slot and the (k, k)-slot for k ̸= i, j and having 0 elsewhere is in N and
satisfies (6.2.13).
316 6. Characteristic Classes
Now consider the elements f˜1 , . . ., f˜n in I(U (n)) corresponding to the
polynomials f1 , . . . , fn on u(n) defined by (6.2.6). The polynomials on G ′
corresponding to Rf˜1 , . . . , Rf˜n are simply the restrictions to G ′ of f1 , . . . , fn
and these are given by
( )k
i
fk (diag(i ξ 1 , . . . , i ξ n )) = Sk (i ξ 1 , . . . , i ξ n )
2π
(−1)k
= Sk (ξ 1 , . . . , ξ n )
(2π)k
is an isomorphism of algebras.
Notice that we have actually proved more. Since the restrictions of f1 , . . . , fn
to G ′ are (up to constants) the elementary symmetric polynomials in ξ 1 , . . . , ξ n ,
they are algebraically independent and therefore the same is true of the ele-
ments Rf˜1 , . . . , Rf˜n in I(G′ ). Thus, f˜1 , . . . , f˜n are algebraically independent
in I(U (n)). Furthermore, Rf˜1 , . . . , Rf˜n generate the algebra IN (G′ ) and so
f˜1 , . . . , f˜n generate I(U (n)) and we have established our major result.
Theorem 6.2.1 Let f˜1 , . . . , f˜n be the elements of I(U (n)) corresponding by
polarization to the ad-invariant polynomials on u(n) defined by (6.2.4) (or
(6.2.6)). Then f˜1 , . . . , f˜n are algebraically independent and generate the alge-
bra I(U (n)).
The result corresponding to Theorem 6.2.1 for SU (n) is almost the same
except that, since su(n) consists of tracefree matrices, f1 is identically zero
and so does not appear in the generating set.
Theorem 6.2.2 Let f˜2 , . . . , f˜n be the elements of I(SU(n)) corresponding
by polarization to the ad-invariant polynomials on su(n) defined by (6.2.4)
(or (6.2.6)). Then f˜2 , . . . , f˜n are algebraically independent and generate the
algebra I(SU (n)).
Exercise 6.2.8 Prove Theorem 6.2.2 by making whatever modifications are
required in the arguments that led to Theorem 6.2.1.
6.3. The Chern-Weil Homomorphism 317
Keep in mind that the generators of I(U (n)) and I(SU (n)) described in
Theorems 6.2.1 and 6.2.2 are actually real-valued on u(n) and su(n), respec-
tively. This will be of some interest when we use them to define the Chern
characteristic classes in the next section.
(6.3.1)
( ))
. . . , Ω p v σ(2k−1) , v σ(2k)
1 ∑ ( )
= k
(−1)σ Ω α
p
1
v σ(1) , v σ(2)
2
σ∈S2k
( ) ( )
. . . Ωα
p
k
v σ(2k−1) , v σ(2k) f˜ eα1 , . . . , eαk .
318 6. Characteristic Classes
But
1 ∑ ( )
(Ω α1 ∧ · · · ∧ Ω αk )p (v 1 , . . . , v 2k ) = (−1)σ Ω α 1
v σ(1) , v σ(2)
2! 2! · · · 2! p
σ∈S2k
( )
αk
. . . Ω p v σ(2k−1) , v σ(2k)
1 ∑ ( )
= k (−1)σ Ω α p
1
v σ(1) , v σ(2)
2
σ∈S2k
( )
. . . Ωα p
k
v σ(2k−1) , v σ(2k)
1[ ]
= f˜ (Ω p (v 1 , v 2 )) − f˜ (Ω p (v 2 , v 1 ))
2
1[ ]
= trace(Ω p (v 1 , v 2 )) − trace(Ω p (v 2 , v 1 ))
2
1
= trace(Ω p (v 1 , v 2 ) − Ω p (v 2 , v 1 ))
2
1
= trace(2Ω p (v 1 , v 2 ))
2
= trace(Ω p (v 1 , v 2 ))
= (trace ◦ Ω)p (v 1 , v 2 )
When
∑ k = 2 the symmetrized trace is given by f˜(A1 , A2 ) = symtr(A1 , A2 ) =
1 1
2! σ∈S2 trace(Aσ(1) Aσ(2) ) = 2 [trace(A1 A2 )+ trace(A2 A1 )] = trace(A1 A2 ).
6.3. The Chern-Weil Homomorphism 319
Thus,
( ) 1 ∑ ( ( )
f˜(Ω) (v 1 , v 2 , v 3 , v 4 ) = 2 (−1)σ f˜ Ω p v σ(1) , v σ(2) ,
p 2
σ∈S4
( ))
Ω p v σ(3) , v σ(4)
1 ∑ ( ( )
= (−1)σ trace Ω p v σ(1) , v σ(2) .
4
σ∈S4
( ))
Ω p v σ(3) , v σ(4)
( 1 ∑ )
= trace (−1)σ Ω p (v σ(1) , v σ(2) .
2! 2!
σ∈S4
( ))
Ω p v σ(3) , v σ(4)
( )
= trace (Ω ∧ Ω)p (v 1 , v 2 , v 3 , v 4 )
= (trace ◦ (Ω ∧ Ω))p (v 1 , v 2 , v 3 , v 4 )
f˜(Ω) = trace(Ω ∧ Ω)
( ) (6.3.4)
f˜(A1 , A2 ) = symtr(A1 , A2 ) = trace(A1 A2 ) .
Exercise 6.3.1 Show that the forms trace Ω and trace(Ω ∧ Ω) on P both
project to closed forms on X. Hint: Consider trace F and
trace(F ∧ F ), where F = s∗ Ω is any field strength corresponding to a local
cross-section s.
Our first major result is a generalization of Exercise 6.3.1.
Theorem 6.3.1 Let ω be a connection with curvature Ω on the principal
P
G-bundle G ,→ P −→ X and let f˜ ∈ I k (G) for some k ≥ 1. Then the 2k-form
f˜(Ω) defined by (6.3.1) projects to a unique closed 2k-form f˜(Ω) on X (i.e.,
f˜(Ω) = P ∗ (f¯(Ω)) and d(f¯(Ω)) = 0).
Proof: To show that f˜(Ω) projects to a unique 2k-form on X we use
Lemma 4.5.1. Thus, we must show
The first of these is clear from (6.3.2) since Ω vanishes when either of its
arguments is vertical. To prove the second we recall that σg∗ (Ω) = adg−1 ◦ Ω
320 6. Characteristic Classes
(Lemma 6.2.2, [N4]) and that f˜ satisfies f˜(adg−1 (A1 ), . . . , adg−1 (Ak )) =
f˜(A1 , . . . , Ak ) so
( ( ))
σg∗ f˜(Ω) (v 1 , . . . , v 2k )
p
( )
= f˜(Ω) ((σg )∗p (v 1 ), . . . , (σg )∗p (v 2k ))
p·g
1 ∑ ( ( )
= k (−1)τ f˜ Ω p·g (σg )∗p (v τ (1) ), (σg )∗p (v τ (2) ) , . . . ,
2
τ ∈S2k
( ))
Ω p·g (σg )∗p (v τ (2k−1) ), (σg )∗p (v τ (2k) )
1 ∑ (
= k (−1)τ f˜ (σg∗ Ω)p (v τ (1) , v τ (2) ), . . . ,
2
τ ∈S2k
)
(σg∗ Ω)p (v τ (2k−1) , v τ (2k) )
1 ∑ ( ( )
τ ˜
= (−1) f ad g −1 Ω p (v τ (1) , v τ (2) ) , . . . ,
2k
τ ∈S2k
( ))
adg−1 Ω p (v τ (2k−1) , v τ (2k) )
1 ∑ ( )
= k (−1)τ f˜ Ω p (v τ (1) , v τ (2) ), . . . , Ω p (v τ (2k−1) , v τ (2k) )
2
τ ∈S2k
( )
= f˜(Ω) (v 1 , . . . , v 2k )
p
as required.
Thus, there exists a unique 2k-form f¯(Ω) on X with P ∗ (f¯(Ω)) = f˜(Ω). We
show that f¯(Ω) is closed. Notice that d(f˜(Ω)) = d(P ∗ (f¯(Ω))) = P ∗ (d(f¯(Ω)))
so d(f˜(Ω)) projects to d(f¯(Ω)). Since projections are unique when they ex-
ist, it will suffice to show that f˜(Ω) is closed. Now, because f˜(Ω) projects,
Lemma 4.5.5 implies that d(f˜(Ω)) = dω (f˜(Ω)) so d(f˜(Ω)) vanishes when any
one of its arguments is vertical. Thus, we need only show that d(f˜(Ω)) van-
ishes when all of its arguments are horizontal. For this we let {eα } be a basis
for G and write Ω = Ω α eα , where each Ω α is a real-valued 2-form on P . Then
so
dω Ω = 0
dω (Ω α eα ) = 0
(dω Ω α )eα = 0
= trace((s∗ Ω)x (w 1 , w 2 ))
= trace(F x (w 1 , w 2 ))
= (trace F )x (w 1 , w 2 ),
1[ ]
= f˜(A1 )f˜(A2 ) + f˜(A2 )f˜(A1 )
2
= f˜(A1 )f˜(A2 )
= (trace A1 )(trace A2 ).
[ω t , ω t ] = [ω 0 + tα, ω 0 + tα]
= [ω 0 , ω 0 ] + t[α, ω 0 ] + t[ω 0 , α] + t2 [α, α]
so that
d
[ω t , ω t ] = [α, ω 0 ] + [ω 0 , α] + 2t[α, α]
dt
= ([ω 0 , α] + t[α, α]) + ([α, ω 0 ] + t[α, α])
= [ω 0 + tα, α ] + [α, ω 0 + tα]
= [ω t , α] + [α, ω t ]
= 2[α, ω t ].
Thus,
( )
d d 1
Ωt = dω t + [ω t , ω t ]
dt dt 2
( )
d 1
= dω 0 + t dα + [ω t , ω t ]
dt 2
d
Ω t = dα + [α, ω t ]. (6.3.9)
dt
then ( )
f˜(φ1 , . . . , φk ) = f˜ eα1 , . . . , eαk φα
1 ∧ · · · ∧ φk .
1 αk
(6.3.11)
First note that this will complete the proof of the Theorem since, if
Φ = P ∗ φ it gives
( ) ( )
d(P ∗ φ) = P ∗ f¯(Ω 1 ) − P ∗ f¯(Ω 0 )
( )
P ∗ (d φ) = P ∗ f¯(Ω 1 ) − f¯(Ω 0 )
Thus,
∫ 1 ( ) ∫ 1
˜ d ˜
dΦ = k d f (α, Ω t , . . . , Ω t ) dt = f (Ω t , Ω t , . . . , Ω t )dt
0 0 dt
= f˜(Ω 1 , Ω 1 , . . . , Ω 1 ) − f˜(Ω 0 , Ω 0 , . . . , Ω 0 ) = f˜(Ω 1 ) − f˜(Ω 0 )
as required.
R (X; C)
∗
w : I(G) −→ Hde
which assigns to each f˜ ∈ I(G) the cohomology class [f¯(Ω)], where Ω is the
curvature of any connection on the bundle. This map is, in fact, an algebra
homomorphism since it is clearly linear and, by Exercise 6.3.2, it also satisfies
( )
w f˜ ⊙ g̃ = w(f˜) ∧ w(g̃).
i
c1 (P ) = [trace F ]. (6.3.17)
2π
326 6. Characteristic Classes
Similarly, from (6.2.9), (6.3.7) and (6.3.8), the 2nd Chern class is
1
c2 (P ) = − [(trace F ) ∧ (trace F ) − trace(F ∧ F )]. (6.3.18)
8π 2
There are similar formulas for the remaining ck (P ). However, since f¯k (Ω) is
a 2k-form on X these will reduce to zero when k > 12 dim X. In particular, if
dim X = 2, then c1 (P ) is the only nonzero Chern class. If dim X = 4, then
c1 (P ) and c2 (P ) are the only nontrivial Chern classes.
Let us consider in somewhat more detail the cases G = U (n) and
G = SU (n) that most interest us. We have already seen that, for these groups,
the Chern classes are real cohomology classes.
Remark: Although we will not do so, one can prove that this is actually
true for any matrix Lie group G; Chern classes are real cohomology classes.
Indeed, they are (in a sense we will describe shortly) integral cohomology
classes.
Furthermore, Theorems 6.2.1 and 6.2.2 imply that the algebra of characteristic
classes for any U (n)- or SU (n)-bundle is generated by the Chern classes.
Remark: This is not true for arbitary Lie groups. However, it is known
that, for any compact, semi-simple Lie group G, the algebra I(G) is finitely
generated so that the Chern-Weil homomorphism for any G-bundle is deter-
mined by its values on finitely many elements of I(G). The images of these
then generate the algebra of characteristic classes. For many of the most im-
portant Lie groups it is possible to write out simple, finite generating sets (see
Chapter XII of [KN2]).
In the case of SU (n), trace F is zero so many of the formulas simplify consid-
erably, e.g.,
c1 (P ) = 0 (G = SU (n)), (6.3.19)
1
c2 (P ) = [trace(F ∧ F )] (G = SU (n)). (6.3.20)
8π 2
We saw in Section 6.1 that U (1)-bundles over S 2 are classified up to equiv-
alence by their 1st Chern classes (note that, for U (1)-bundles, F is a 1 × 1
matrix so trace F is just its sole entry and the definition of c1 (P ) given in
Section 6.1 agrees with (6.3.17)). We would like to show that SU (2)-bundles
over S 4 are likewise characterized by their 2nd Chern classes. First we show
that equivalent bundles have, not only the same Chern classes, but the same
Chern-Weil homomorphisms.
P P′
Theorem 6.3.3 Let G ,→ P −→ X and G ,→ P ′ −→ X be equivalent
principal G-bundles over X. Then the Chern-Weil homomorphisms
I(G) −→ H ∗ (X; C ) they determine are equal.
6.4. Chern Numbers 327
as required.
Exercise 6.3.6 Show that the algebra of characteristic classes for a trivial
R (X; C). Hint: The product bundle
∗
bundle is the trivial subalgebra of Hde
π
G ,→ X × G −→ X admits a flat connection (page 353, [N4]).
and we wish to apply Stokes’ Theorem to each of the integrals on the right-
hand side. To do so, of course, we need to express both trace(F N ∧ F N ) and
trace(F S ∧ F S ) as exterior derivatives of 3-forms on UN and US , respectively.
This is possible, of course, since they are closed forms on discs, but we will
require more explicit information.
P
Lemma 6.4.1 Let SU (2) ,→ P −→ X be a principal SU (2)-bundle over X, ω
a connection on it with curvature Ω, s : V −→ P −1 (V ) a local cross-section
and A = s∗ ω and F = s∗ Ω the local gauge potential and field strength. Then,
on V ,
( ( ))
2
trace(F ∧ F ) = d trace A ∧ dA + A ∧ A ∧ A . (6.4.3)
3
Proof: Throughout the proof we will work exclusively on the trivial bundle
SU (2) ,→ P −1 (V ) −→ V , but, for convenience, we will denote any relevant
restrictions by the same symbo1, e.g., P is the projection on P −1 (V ), ω is
P
the connection on SU (2) ,→ P −1 (V ) −→ V , etc. We intend to apply (6.3.14),
where f˜ is f˜2 , Ω 1 is Ω and Ω 0 is the curvature of a connection whose existence
we now ask the reader to demonstrate.
P
Exercise 6.4.2 Let G ,→ P −→ X be any principal G-bundle,
s : V −→ P −1 (V ) a local cross-section and Ψ : P −1 (V ) −→ V × G the asso-
ciated trivialization (so s(x) = Ψ−1 (x, e) for every x ∈ V ). Show that there
exists a connection ω 0 on G ,→ P −1 (V ) −→ V for which the local gauge po-
tential A0 = s∗ ω 0 is identically zero. Hint: If Θ is the Cartan 1-form for G
and π : V × G −→ G is the projection, then π ∗ Θ is a flat connection on V × G
(page 353, [N4]). Show that ω 0 = (π ◦ Ψ)∗ Θ has the required properties.
Now, (6.3.14) asserts that
where
∫ 1
Φ=2 f˜2 (α, Ω t )dt
0
α = ω − ω0
ω t = ω 0 + tα
1
Ω t = dωt ω t = dω 0 + t dα + [ω t , ω t ]
2
= dω 0 + t dα + ω t ∧ ω t .
But
α ∧ Ω t = α ∧ (dω 0 + t dα + ω t ∧ ω t )
= α ∧ dω 0 + tα ∧ dα + α ∧ (ω 0 + tα) ∧ (ω 0 + tα)
= α ∧ dω 0 + tα ∧ dα + α ∧ ω 0 ∧ ω 0 + tα ∧ ω 0 ∧ α
+ tα ∧ α ∧ ω 0 + t2 α ∧ α ∧ α
so
∫ 1 ∫ 1
Φ=2 f˜2 (α, Ω t )dt = 2 trace(α ∧ Ω t )dt
0 0
( ∫ 1 )
= trace 2 α ∧ Ω t dt
0
(
= trace 2α ∧ dω 0 + α ∧ dα + 2α ∧ ω 0 ∧ ω 0 + α ∧ ω 0 ∧ α
)
2
+α ∧ α ∧ ω 0 + α ∧ α ∧ α .
3
Notice that, as expected, the integral in (6.4.7), i.e., the Chern number,
does not depend on the connection or the curvature, but only on the transition
function gSN of the bundle. To obtain a computationally more efficient formula
−1 ∗
we expand gSN dgSN = gSN Θ in terms of the basis
( ) ( ) ( )
i 0 0 1 0 i
I = , J = , K =
0 −i −1 0 i 0
for su(2) ∼
= Im H. Thus, let
( )
−1 1 2 3 Θ1 i Θ2 + Θ3 i
gSN dgSN = Θ I + Θ J + Θ K = .
−Θ 2 + Θ 3 i − Θ1 i
332 6. Characteristic Classes
−1 −1 −1
trace((gSN dgSN ) ∧ (gSN dgSN ) ∧ (gSN dgSN )) = −12Θ 1 ∧ Θ 2 ∧ Θ 3 .
Thus,
∫ ∫
1 1
trace(F ∧ F ) = ι∗ (Θ 1 ∧ Θ 2 ∧ Θ 3 ). (6.4.8)
8 π2 S4 2π 2
S3
Our objective now is to show that the Chern number captures the topolog-
ical type of an SU (2)-bundle over S 4 . For this we recall that the equivalence
class of such a bundle is uniquely determined by the homotopy type of gSN | S 3
(this is essentially the Classification Theorem for bundles over spheres, but
Lemma 4.4.1 and Theorem 4.4.2 of [N4] make the correspondence explicit).
Our procedure then will be to select one representative from each homotopy
class in π3 (SU (2)) ∼ = π3 (S 3 ) and show that the bundles with these transition
functions have different Chern numbers.
The notation is less cumbersome if we identify both S 3 and SU (2) with the
unit quaternions. Fix a base point 1 ∈ S 3 and identify π3 (S 3 ) with the set
[(S 3 , 1), (S 3 , 1)] of homotopy classes of maps g : (S 3 , 1) −→ (S 3 , 1) (see page
154, [N4]). For each n = 0, 1, 2, . . . we let
gn : (S 3 , 1) −→ (S 3 , 1)
Show also that, on S 3 ⊆ H, the real part of q −1 dq is zero and the wedge
product of the i , j and k components is
6.4. Chern Numbers 333
1 ∗( −1 −1 −1
)
ω=− ι trace((q dq) ∧ (q dq) ∧ (q dq))
24π 2
so
1 ( )
g∗ ω = − 2
(ι ◦ g)∗ trace((q −1 dq) ∧ (q −1 dq) ∧ (q −1 dq))
24π
1 (
=− 2
trace (ι ◦ g)−1 d(ι ◦ g) ∧ (ι ◦ g)−1 d(ι ◦ g)
24π )
∧ (ι ◦ g)−1 d(ι ◦ g) .
Show that
c(P ) = 1 + c1 (P ) + c2 (P ) + · · ·
of all the Chern classes (necessarily finite since ck (P ) = 0 when 2k > dim X)
is called the total Chern class of the bundle. For bundles with other struc-
ture groups one can often isolate analogous finite generating sets and build
6.5. Z2 -Čech Cohomology for Smooth Manifolds 335
L+↑ ,→ X × L+↑ −→ X.
In this case one can build a spinor structure for X from the product bundle
SL(2, C) ,→ X × SL(2, C) −→ X
by simply defining λ to be the spinor map Spin on each fiber {x} × SL(2, C).
Indeed, X need not be trivial for this to work. All that’s required is that
the frame bundle be trivial and this is the case whenever it has a global
336 6. Characteristic Classes
Suppose now that the frame bundle L(X) is not trivial. It will simplify
the discussion and not obscure any of the essential ideas if we assume that
X is simply connected. In this case it can occur that the frame bundle L(X)
is also simply connected (this cannot occur when the frame bundle is trivial
since π1 (X × L+↑ ) ∼
= π1 (X) ⊕ π1 (L+ ) ∼
↑
= π1 (X) ⊕ Z2 which is just Z2 when X
is simply connected). Now, fix an x0 ∈ X and consider the fiber PL−1 (x0 ) ∼ =
↑
L+ above x0 . In Section 3.4 we constructed a curve R1 (t), 0 ≤ t ≤ 2π, in
L+↑ representing a continuous rotation of the spatial coordinate axes through
360o . We now regard R1 as a curve in the fiber PL−1 (x0 ). Within PL−1 (x0 ), R1
is not nullhomotopic. However, if L(X) is simply connected, the loop R1 (t),
0 ≤ t ≤ 2π, is nullhomotopic in L(X) (see the figure).
–1
(x0) ≅ +
R1
(X)
X x0
But a Dirac spinor field (for example) must change sign under( a )360◦
rotation at x0 (because R1 lifts to a path in SL(2, C) from 10 01 to
( )
− 10 01 ) and obviously does not change sign under “no rotation at all.” Since
the wavefunction cannot change continuously from ψ to −ψ but would have to
do so over the homotopy in L(X) from R1 to the trivial loop, it is not possible
to unambiguously define such spinor fields on X. With this as motivation we
now set about finding a characteristic class, the vanishing of which prohibits
this sort of topological obstruction.
Recall (Section 3.3) that an open cover {Uα }α∈A of a manifold X is said
to be simple if any finite intersection Uα0 ∩ · · · ∩ Uαj of its elements is ei-
ther empty or diffeomorphic to Rn (where n = dim X). Smooth manifolds
6.5. Z2 -Čech Cohomology for Smooth Manifolds 337
admit simple covers and, indeed, any open cover of a smooth manifold has
a countable, locally finite, simple open refinement in which each element has
compact closure. Our construction of the Čech cohomology groups begins with
the selection of a locally finite, simple open cover U = {Uk }k=1,2,... for X with
each Ūk compact. We first build the Čech cohomology groups Ȟ j (U; Z2 ) cor-
responding to this cover and then discuss how one goes about showing that
the construction is actually independent of this choice. We will identify Z2
with the multiplicative group {−1, 1}.
For each integer j ≥ 0, a j -simplex (for U) is an ordered (j + 1)-tuple
σ = (k0 , . . . , kj ) of indices for which Uk0 ∩ · · · ∩ Ukj is nonempty (and
therefore diffeomorphic to Rn ). The support of σ is the nonempty open
set |σ| = Uk0 ∩ · · · ∩ Ukj . For i = 0, . . . , j, the i th -face of the j-simplex
σ = (k0 , . . . , kj ) is the (j − 1)-simplex σ i = (k0 , . . . , ki−1 , ki+1 , . . . , kj ). The
set of all j-simplexes for j ≥ 0 is called the nerve of U and denoted N (U).
For each j ≥ 0, a (Z2 -Čech) j -cochain is a function f which assigns to each
j-simplex σ = (k0 , . . . , kj ) an element f (σ) = f (k0 , . . . , kj ) of Z2 and that is
totally symmetric, i.e.,
( )
f kτ (0) , . . . , kτ (j) = f (k0 , . . . , kj )
for all permutations τ ∈ Sj+1 . The set Č j (U; Z2 ) of all j-cochains is an Abelian
group under pointwise multiplication, i.e.,
for all j-simplexes (k0 , . . . , kj ). Notice that the identity element of this group
assigns the value 1 ∈ Z2 to each j-simplex and that every element of Č j (U; Z2 )
is its own inverse.
Before proceeding with the construction we would like to persuade the
reader that Čech cochains actually arise in nature, as it were. Suppose
then that X is a semi-Riemannian manifold and consider the orthonor-
P
mal frame bundle O(m, n − m) ,→ F (X) −→ X. Select a simple cover
U = {Uk }k=1,2,... for X of the sort described above. Since each Uk is dif-
feomorphic to Rn , the bundle is trivial over Uk and so we may select an
orthonormal frame field on Uk , i.e., a cross-section sk : Uk −→ F (X).
Do this for each element of U. Now, if k0 and k1 are two indices for
which Uk0 ∩ Uk1 ̸= ∅, then sk0 and sk1 are related on Uk0 ∩ Uk1 by
sk1 = sk0 · gk0 k1 , where gk0 k1 is the transition function. Since this tran-
sition function takes values in O(m, n − m), its determinant is either 1
or −1 at each point Uk0 ∩ Uk1 . But Uk0 ∩ Uk1 is diffeomorphic to Rn
(and therefore connected) so det(gk0 k1 ) is either 1 everywhere on Uk0 ∩ Uk1
or −1 everywhere on Uk0 ∩ Uk1 . Thus, we may define f (k0 , k1 ) ∈ Z2
by ( )
f (k0 , k1 ) = det gk0 k1 .
338 6. Characteristic Classes
f (k1 , k0 ) = f (k0 , k1 )
and
∏
j+2 ( )
(δ j+1 (δ j f ))(k0 , . . . , kj+2 ) = (δ j f ) k0 , . . . , k̂i , . . . , kj+2
i=0
= (δ j f )(k1 , k2 , . . . , kj+2 )(δ j f )(k0 , k2 , . . . , kj+2 )
· · · (δ j f )(k0 , k1 , . . . , kj+1 )
[ ]
= f (k2 , k3 , . . . , kj+2 )f (k1 , k3 , . . . , kj+2 ) · · · f (k1 , k2 , . . . , kj+1 )
[ ]
· f (k2 , k3 , . . . , kj+2 )f (k0 , k3 , . . . , kj+2 ) · · · f (k0 , k2 , . . . , kj+1 ) · · · ·
[ ]
· f (k1 , k2 , . . . , kj+1 )f (k0 , k2 , . . . , kj+1 ) · · · f (k0 , k1 , . . . , kj ) .
Each of these factors has precisely two of the indices k0 , . . . , kj+2 missing.
Moreover, for any pair of indices, there are precisely two factors in which
these indices are missing. Thus, any factor that takes the value −1 occurs
precisely twice and the product must be 1.
Lemma 6.5.1 asserts that the sequence of cochain groups and coboundary
operators forms a cochain complex (Section 5.3) and so we may build its
cohomology theory in the usual way. A Čech j -coboundary is an element
of Č j (U; Z2 ) that is in the image of δ j−1 and the set
Ž j (U; Z2 ) = ker δ j
B̌ j (U; Z2 ) ⊆ Ž j (U; Z2 )
f ′ = (δ j−1 h)f
=1
[f ] ∈ Ȟ 1 (U; Z2 ).
One disturbing aspect of this last example is that the cohomology class [f ]
is not obviously intrinsic to the topology of X as is the case, for example,
for de Rham cohomology classes. Indeed, it appears to depend not only on
the simple cover U from which it is constructed, but even on the particu-
lar orthonormal frame fields sk : Uk −→ F (X) selected on each of the el-
ements of U. Let us show that at least this last dependence on the cross-
sections sk is only apparent. Thus, we let s′k : Uk −→ F (X) be another
orthonormal frame field on Uk for each k = 1, 2, . . . . Both sk and s′k give
rise to trivializations of the bundle defined on P −1 (Uk ) and these trivializa-
tions are related by a transition function gk defined on Uk (so s′k = sk · gk
on Uk ).
Exercise 6.5.2 Let (k0 , k1 ) be a 1-simplex with sk1 = sk0 · gk0 k1 and s′k1 =
s′k0 · gk′ 0 k1 on Uk0 ∩ Uk1 . Show that gk′ 0 k1 = gk−1
0
gk0 k1 gk1 on Uk0 ∩ Uk1 .
Now, defining a 1-cocycle f ′ just as we did for f , but using the cross-section
s′ rather than s gives
( ) ( )
f ′ (k0 , k1 ) = det gk0 k′ = det gk−1
0
gk0 k1 gk1
1
( ( ) ( ))
= det gk0 det gk1 f (k0 , k1 ).
i.e.,
f ′ = (δ 0 h)f
ρj : Č j (U; Z2 ) −→ Č j (V; Z2 )
by
( )
(ρj f )(l0 , . . . , lj ) = f ρ(l0 ), . . . , ρ(lj )
for each f ∈ Č j (V; Z2 ) and every j-simplex (l0 , . . . , lj ) for V. For j < 0 we
take ρj : Č j (U; Z2 ) −→ Č j (V; Z2 ) to be the trivial homomorphism.
Lemma 6.5.2 For each integer j, ρj+1 ◦ δ j = δ j ◦ ρj .
342 6. Characteristic Classes
δj
2
ρj ρ j +1
Čj ( ; 2) Č j+1 ( ; 2)
δj
∏
j+1 ( )
= f ρ(l0 ), . . . , ρ(lc
i ), . . . , ρ(lj+1 )
i=0
( )
= (δ j f ) ρ(l0 ), . . . , ρ(lj+1 )
∏
j−1 ( )
j
(K f )(l0 , . . . , lj−1 ) = f ρ(l0 ), . . . , ρ(li ), τ (li ), . . . , τ (lj−1 )
i=0
and conclude that the cochain maps {ρj } and {τ j } are algebraically homotopic
and therefore induce the same maps in cohomology.
The next step (which we do not intend to take) would be to show that
the homomorphisms (ρj )# are, in fact, isomorphisms (again, we refer to
[GR]). Granting this, we will henceforth call the groups we have constructed
the Z2 - Čech cohomology groups of X and denote them Ȟ j (X; Z2 ). Any
calculation of these groups for some specific manifold X would, of course,
begin with the selection of some convenient simple cover. For example,
Exercise 6.5.4 now implies that Ȟ 0 (Rn ; Z2 ) ∼ = Z2 and Ȟ j (Rn ; Z2 ) is
trivial for every other j. Our particular concern will be with isolating cer-
tain specific cohomology classes which act as “obstructions” to the exis-
tence of various desirable structures on a manifold. For example, the coho-
mology class in Ȟ 1 (X; Z2 ) associated with the orthonormal frame bundle
P
O(m, n − m) ,→ F (X) −→ X of a semi-Riemannian manifold X in the man-
ner described above is called the 1st Stiefel-Whitney class of X and is
denoted w1 (X). The next order of business is to show that the 1st Stiefel-
Whitney class w1 (X) ∈ Ȟ 1 (X; Z2 ) is the obstruction to orientability for a
semi-Riemannian manifold X.
so
( ) ( ) ( ) ( )
det gk′ 0 k1 = det gk0 det gk1 det gk0 k1
= f0 (k0 ) f0 (k1 ) f (k0 , k1 )
= (δ0 f0 ) (k0 , k1 ) f (k0 , k1 )
= (f (k0 , k1 ))2
= 1.
Since s′k1 = s′k0 · gk′ 0 k1 for every 1-simplex (k0 , k1 ), the orthonormal frames
s′k0 and s′k1 therefore determine the same orientation at each p ∈ Uk0 ∩ Uk1 .
Consequently, the collection of all s′k : Uk −→ F (X), k = 1, 2, . . . , determines
an orientation on X.
λ X
+ (X)
PL ◦ λ = PS
and
λ(p · g) = λ(p) · Spin (g)
λ × Spin λ X
(X)× + (X)
6.5. Z2 -Čech Cohomology for Smooth Manifolds 345
Let U = {Uk }k=1,2,... be a locally finite, simple open cover of X with each
Ūk compact and choose local orthonormal frame fields sk : Uk −→ L(X),
k = 1, 2, . . . . The corresponding transition functions
map into L+↑ for every 1-simplex (k0 , k1 ). Because each Uk0 ∩ Uk1 is diffeomor-
phic to R4 and Spin : SL(2, C) −→ L+↑ is a covering map, these transition
functions lift to SL(2, C):
SL ( 2, C )
k0 k1 Spin
U k ∩ Uk +
0 1
k0 k1
g̃k1 k0 = g̃k0 k1 −1
at each point in Uk0 ∩ Uk1 ∩ Uk2 . But this intersection is connected (diffeo-
morphic to R4 ) and so this map is either id everywhere or −id everywhere.
Thus, we can define a map Z from 2-simplexes to Z2 by
Z ′ = (δ 1 h)Z
gk Spin
Uk gk +
by
hk0 k1 = g̃k0 k1 g̃k1 g̃k′ 1 k0 g̃k−1
0
.
Since Spin is a homomorphism, Spin ◦ hk0 k1 takes the value id ∈ L+↑ for each
x in Uk0 ∩ Uk1 . Thus, hk0 k1 maps into {±id} ⊆ SL(2, C) and connectivity of
Uk0 ∩ Uk1 implies that it must be constant. Thus, we may define a 1-cochain
h by
h(k0 , k) == hk0 k1 (x)
by
g̃k′ 0 k1 = g̃k0 k1 f (k0 , k1 ).
Then
so the primed lifts satisfy (6.5.1). Now we will prove our major result by show-
ing that the existence of lifts satisfying (6.5.1) is equivalent to the existence
of a spinor structure on X.
Proof: For both parts of the proof we fix at the outset a locally finite, simple
open cover U = {Uk }k=1,2,... for X.
Suppose first that a spinor structure
λ × Spin λ X
(X)× + (X)
P
for X exists. The spinor bundle SL(2, C) ,→ S(X) −→ S
X is trivial over each
Uk so we can choose a cross-section
( )
s̃k : Uk −→ PS−1 (Uk ) = λ−1 PL−1 (Uk )
sk : Uk −→ PS−1 (Uk )
by
sk = λ ◦ s̃k .
348 6. Characteristic Classes
P
Then each sk is a cross-section of the frame bundle L+↑ ,→ L(X) −→ L
X
because PL ◦ λ = Ps . These cross-sections give rise to trivializations and
therefore transition functions and we denote the transition functions by g̃k0 k1
and gk0 k1 , respectively.
k0 k1 Spin
U k ∩ Uk +
0 1
k0 k1
with the g̃k0 k1 as transition functions. All that remains is to define a map
λ : S(X) −→ L(X)
satisfying PL ◦ λ = PS and λ(p · g) = λ(p) · Spin (g). For this we will define
λ above each Uk and then show that the definitions agree on any nonempty
intersections.
The trivialization Ψk : PL−1 (Uk ) −→ Uk × L+↑ determined by sk is given by
for all x ∈ Uk and g ∈ L+↑ . Let Ψ̃k : PS−1 (Uk ) −→ Uk × SL(2, C),
k = 1, 2, . . . , be trivializations of S(X) related by the transition functions
g̃k0 k1 and let s̃k : Uk −→ PS−1 (Uk ), k = 1, 2, . . . , be the corresponding canon-
ical cross-sections. Define
by
λk = Ψ−1
k ◦ (idUk × Spin) ◦ Ψ̃k
for each k = 1, 2, . . . .
Exercise 6.5.9 Show that λk (s̃k (x) · g) = sk (x) · Spin(g) for all x ∈ Uk and
all g ∈ SL(2, C) and conclude that, on PS−1 (Uk ),
PL ◦ λk = PS
and
λk (p · g) = λk (p) · Spin (g).
and
sk1 (x) = sk0 (x) · gk0 k1 (x)
so
( ) ( ( ))
λk0 s̃k1 (x) · g = λk0 s̃k0 (x) · g̃k0 k1 (x)g
( )
= sk0 (x) · Spin g̃k0 k1 (x)g
( ( ))
= sk0 (x) · Spin g̃k0 k1 (x) · Spin(g)
( )
= sk0 (x) · gk0 k1 (x) · Spin(g)
as required.
350 6. Characteristic Classes
Theorem 6.5.4 has some rather obvious consequences that are of sufficient
importance to be recorded officially. If X happens to be diffeomorphic to R4 ,
then Ȟ 2 (X; Z2 ) is trivial so w2 (X) is trivial and X admits a spinor structure.
Corollary 6.5.5 A time oriented spacetime diffeomorphic to R4 admits a
spinor structure.
This applies, in particular, to Minkowski spacetime and the Einstein-deSitter
spacetime. The deSitter spacetime and the Einstein cylinder are diffeomorphic,
not to R4 , but to S 3 × R. That they nevertheless do admit spinor structures
will follow from our next result.
Corollary 6.5.6 Let X be an oriented, time oriented spacetime and suppose
that the oriented, time oriented orthonormal frame bundle
↑ PL
L+ ,→ L(X) −→ X is trivial. Then X admits a spinor structure.
Proof: By assumption, there exists a global cross-section s : X −→ L(X).
Letting U = {Uk }k=1,2,... be a locally finite simple open cover we define sk :
Uk −→ PL−1 (Uk ) by sk = s|Uk . Since sk0 = sk1 · gk0 k1 on Uk0 ∩ Uk1 , all of
the corresponding transition functions gk0 k1 are identically equal to id ∈ L+↑ .
These surely lift to maps g̃k0 k1 : Uk0 ∩ Uk1 −→ SL(2, C) that take the value
id ∈ SL(2, C) everywhere and these satisfy (6.5.1) so w2 (X) is trivial and X
admits a spinor structure.
P
Remark: Geroch [G1] has shown that the triviality of L+↑ ,→ L(X) −→ L
X
actually characterizes the existence of spinor structures for oriented, time
oriented (noncompact) spacetimes.
Since we have shown in Chapter 3 that both deSitter spacetime and the Ein-
stein cylinder satisfy the hypotheses of Corollary 6.5.6., they too admit spinor
structures.
Appendix
Regrettably, this integer need not be an invariant and may, in fact, depend
on the choice of generic metric g . To prevent this one must ensure that not
only is it possible to choose a metric for which there are no reducible ASD
connections, but also that a generic variation of that metric (generic path in
the space R of Riemannian metrics on M ) does not introduce reducibles. For
this one needs the set of metrics in R for which there are reducibles to be
sufficiently “thin” and this, by the Remark above, means that b+2 (M ) must be
sufficiently large. One can show that
b+
2 (M ) > 1
µ : H2 (M ) −→ H 2 (M(Pk , g ))
b+
2 (M ). Henceforth, we assume
is only in this last step that the values of γd (M ) become multiples of 12 rather
than just integers. Those inclined to find out what all of this really means are
referred to [M2] and [DK].
It is certainly not our intention here to compute Donaldson invariants or use
them to obtain topological information about 4-manifolds. Rather we would
like to sketch how, by adopting a slightly different perspective, they lead, by
way of quantum field theory, to the Seiberg-Witten theory that is our real
concern here. To keep the discussion as uncluttered as possible we will focus
most of our attention on γ0 (M ) and will economize on notation by writing P
for Pk , G for G(Pk ), A for A(Pk ), etc.
The gauge group G does not act freely on the space  of irreducible con-
nections since even irreducibles have a Z2 stabilizer. However, Ĝ = G/Z2 does
act freely on  so we have an infinite-dimensional principal bundle
Ĝ ,→ Â −→ B̂.
We build a vector bundle associated to this principal bundle as follows: We
claim that there is a smooth left action of Ĝ on the (infinite-dimensional)
vector space Λ2+ (M , ad P ) of self-dual 2-forms on M with values in the adjoint
bundle ad P . To see this we think of G as the group of sections of the nonlinear
adjoint bundle Ad P under pointwise multiplication. Since the elements of
Λ2+ (M , ad P ) take values in the su(2)-fibers of ad P, G acts on these fibers by
conjugation. Moreover, conjugation takes the same value at ±f ∈ G so this
G-action on Λ2+ (M , ad P ) descends to a G/Z2 = Ĝ-action. Thus, we have an
associated vector bundle
F + : Â −→ Λ2+ (M, ad P )
F + (ω) = F +
ω =
1
2 (F ω + ∗ F ω ) .
of our vector bundle. Now notice that the moduli space M of anti-self-dual
connections (F +ω = 0) is precisely the zero set of the section s+ (for generic
g all such connections are irreducible). It is a general fact that the base of
any smooth vector bundle is diffeomorphic to the image of any cross-section
of the vector bundle (in particular, the one that picks out the zero element in
each fiber). Thus, we can identify B̂ with the image of the zero section
We conclude that ( ) ( )
M = s+ B̂ ∩ s0 B̂ .
where e(X) is called the Euler class of X. This is a characteristic class that
can be defined by the Chern-Weil procedure described in Section 6.3. Briefly,
the construction goes like this: Choose a Riemannian metric on X and consider
the corresponding oriented, orthonormal frame bundle
SO(2k) ,→ F+ (X) −→ X
P f : so(2k) −→ R.
Although there are much more elegant ways to define this (see [MS]) we will
opt for a simple-minded formula in terms of the matrix entries. Let A = (Aij )
358 Appendix
where S2k is the group of permutations of {1, . . . 2k} and (−1)σ denotes the
sign of the permutation σ. One can check, for example, that if
0 λ
1
0
−λ
1 0
A= . .
.
0 λk
0
−λk 0
then P f (A) = λ1 · · · λk (in fact, it is always the case that (P f (A))2 = det A).
One can show that P f is invariant under the adjoint action of SO(2k) on
so(2k) so Chern-Weil guarantees that if we write Ω = (Ωij ) as a matrix of
1-forms (and normalize by − 2π 1
), then
( )
1 (−1)k ∑
Pf − Ω = (−1)σ Ωσ(1)σ(2) ∧ · · · ∧ Ωσ(2k−1)σ(2k)
2π 22k π k k!
σ∈S2k
descends to a closed 2k-form on X whose cohomology class e(X) does not
depend on the choice of the connection Ω. This is the Euler class and its
integral over X is the Euler characteristic (this is proved in [MT]).
All of this can be generalized to an arbitrary oriented, real vector bundle
E −→ X of fiber dimension 2k over a compact, oriented manifold of dimension
2k. Here one chooses a fiber metric (smoothly varying positive definite inner
products on the fibers of E) to get an oriented, orthonormal frame bundle
SO(2k) ,→ F+ (E) −→ X. The Euler class e(E) is then defined just as above
and one defines the Euler number χ(E) of the bundle to be its integral over X
(this is no longer the Euler characteristic of X, of course). One can then prove
an analogue of the Poincaré-Hopf Theorem that gives χ(E) as the intersection
number for a cross-section of E (see [MT]).
What we have seen is that the Donaldson invariant γ0 (M ) is analogous to
the Poincaré-Hopf version of an Euler number (although the vector bundle
is infinite-dimensional and the Poincaré-Hopf Theorem itself is valid only in
finite dimensions). If one were to take this analogy seriously it might suggest
the possibility of an integral representation of γ0 (M ) analogous to the Gauss-
Bonnet-Chern Theorem. Notice, however, that such an integral would be over
the base of the vector bundle which is the infinite-dimensional moduli space B̂.
Now, integrals over infinite-dimensional manifolds are notoriously difficult to
make rigorous mathematical sense of, but, fortunately, this does not bother the
A.1. Donaldson Invariants and TQFT 359
physicists at all. As Nigel Hitchin has said, “This is such stuff as quantum field
theory is made of.” Indeed, it was Edward Witten [W2] who first produced a
(formal) integral representation of γ0 (M ), not directly, but as what is called
the partition function of a certain variant of supersymmetric quantum Yang-
Mills theory. Indeed, this quantum field theory also yielded formal integral
representations for all of the Donaldson invariants and eventually led to the
Seiberg-Witten invariants. Athough we are not so presumptuous as to attempt
any sort of exegesis of Witten’s work a quick tour of a few of the ideas is the
only way to see the emergence of Seiberg-Witten.
There are various approaches to the construction of a quantum field theory,
but the only one that will concern us here is the so-called Feynmann path
integral approach. Here one begins with a classical field theory of just the
sort we discussed in Chapter 2. Thus, one is given a collection ξ of classical
fields (gauge fields, i.e., connections, and matter fields, i.e., sections of vector
bundles) and an action S(ξ). The action has various symmetries (e.g., gauge
invariance, relativistic invariance, etc.) so that the physically significant object
of study is the moduli space F/S of fields modulo symmetrics. Real-valued
functions
O : F/S −→ R
on the moduli space are called observables. In this context, “quantization” is
viewed as the process of assigning expectation values ⟨ O ⟩ to
observables O and, according to the rules of the game, this is accomplished
by a weighted integral over the moduli space which, when the underlying
manifold is Riemannian (as opposed to semi-Riemannian), is of the form
∫
e−S(ξ)/e O([ξ]) [Dξ].
2
⟨O⟩ =
F/S
4-manifold with b+ 2 (M ) > 1 and odd and we choose some generic metric g
on M . The field content consists of one gauge field (connection) ω and five
matter fields ϕ, λ, η, χ and ψ all of which are forms (of various degrees) with
values in the adjoint bundle ad P . But this gauge theory of Witten’s is super-
symmetric which means that each matter field is classified as either “bosonic”
or “fermionic” and there is an additional (super) symmetry operator that in-
terchanges bosons and fermions. We will make no attempt to explain what
this means, but will simply refer the curious to [W1] and record the types of
the matter fields:
Bosonic Fermionic
ϕ ∈ Λ (M, ad P )
0
η ∈ Λ0 (M, ad P )
λ ∈ Λ0 (M, ad P ) χ ∈ Λ1 (M, ad P )
ψ ∈ Λ2+ (M, ad P )
Remarks: A few of the terms in the action are familiar. The first is a Yang-
Mills term. The second Witten calls a topological term since it is essentially
the Chern class. The rest are to be regarded as interaction terms. The path
Witten followed to arrive at the field content, the symmetries and the terms in
the action is described in detail in [W2], but it is not a path easily traversed
by mathematicians. Remarkably, Atiyah and Jeffrey [AJ] have shown that the
fields and the action all arise naturally from purely geometrical considerations
by formally applying to the infinite-dimensional vector bundle  ×Ĝ Λ2+ (M ,
ad P ) a formula for the Euler class proved (for finite-dimensional vector bun-
dles) by Mathai and Quillen [MQ].
SDW (ξ) is a perfectly well-defined mathematical object, but now we must
quantize and this means integrating over the entire moduli space of field con-
figurations. The partition function, for example, is formally written as the
path integral ∫
e−SDW (ξ)/e [Dξ].
2
ZDW =
A.1. Donaldson Invariants and TQFT 361
Now, a path integral is really not an integral at all, but just a suggestive
notation for a certain limit (a limit that generally does not exist). Physicists
have developed elaborate techniques for dealing with such “integrals”, but
we will say nothing about this. Rather we will simply sketch, in very broad
terms, how Witten was led to identify ZDW with γ0 (M ) in the case in which
8k − 3(1 + b+ 2 (M )) = 0. The crucial observation is that ZDW , whatever it
means, is a function of the coupling constant e. At the classical level, e plays
no real role and can simply be regarded as a rescaling of the action, but in
the quantum theory its size will determine the computability (or not) of the
relevant physical quantities. This is because the usual procedure for dealing
with these quantities is to do perturbation calculations and this involves series
expansions in e. If e is “small” (i.e., in what is called the “weak coupling
limit”) such calculations are extraordinarily effective, but if e is “large” (“strong
coupling”) they fail completely.
Witten computed ZDW is the weak coupling limit by performing what the
physicists call the “semi-classical approximation” which he concluded, based
on all of the symmetries built into the action, must, in fact, be exact.
Remark: This is an infinite-dimensional analogue of the famous
Duistermaat-Heckman Theorem on exact stationary phase approximations
(see [BV]).
From this he argued that the path integral defining ZDW “localizes” to an
integral over the moduli space M of anti-self-dual connections on P and,
when dim M = 0, this integral over the 0-dimensional moduli space is just a
sum that can be identified with the Donaldson invariant γ0 (M ).
Remark: Intuitively, this is not unlike the Residue Theorem which localizes
a contour integral around a closed path to a sum of contributions from the
singularities of the integrand inside (although there are no “symmetry” con-
siderations here). Much closer mathematical analogues are the Equivariant
Localization Theorems of Berline and Vergne (see [BV]) which extract and
generalize the essential content of the Duistermaat-Heckman Theorem.
Witten also isolated observables in his TQFT whose expectation values
formally coincide with the integrals we used in our “naive” definition of
γd (M ) for d > 0. Very briefly, it goes like this: For any field configuration
ξ = (ω, ϕ, λ, η, χ, ψ) define W = trace( 14 ψ ∧ ψ + ϕF ω ). This is an ordinary
2-form on M . For each homology class x ∈ H2 (M ), thought of as an embedded
surface, define ∫
O(x) = W.
x
Each O(x) maps a field configuration to a real number. Again due to all of the
built-in symmetries, O(x) is actually defined on the moduli space and so is
an observable. Witten associates with each O(x) a closed 2-form α(x) on the
moduli space M whose cohomology class can be identified with µ(x)(µ is the
362 Appendix
All of this is quite formal, of course, and the end result is not really the
Donaldson invariants but only our naive description of some of them. Never-
theless, it is remarkable. Still more remarkable is the fact that this is just the
beginning of the story and that the best part is yet to come. To relate this part
of the story we must first say a few words about another feature of Witten’s
TQFT. This is a very subtle and deep type of “duality” that we cannot do
justice to here, but which has its origins in the fact that, unlike most quan-
tum fields theories, its partition function and expectation values are actually
independent of the coupling constant e. This is not a mathematical theorem,
of course, but a consequence of formal path integral manipulations based on
the myriad symmetries of the action. Ordinarily one would expect the weak
and strong coupling regimes to describe very different physical systems, but
in the case at hand they must be entirely equivalent, although their mathe-
matical descriptions would no doubt look quite different (since, for example,
perturbation calculations are possible in one, but not the other). This suggests
the possibility of an entirely different description of the Donaldson invariants
buried in the strong coupling regime. Witten was well aware of this in 1988
when [W2] appeared, but could do nothing about it because no one knew
how to compute anything in the strong coupling (nonperturbative) regime.
And so matters stood until 1994 when Seiberg and Witten developed entirely
new techniques for doing exact calculations in strong coupling for so-called
N = 2 supersymmetric Yang-Mills theories. This done, Witten dusted off his
old TQFT, applied the new techniques and uncovered the dual version of Don-
aldson theory. This is, of course, the Seiberg-Witten theory that we have been
leading up to all along. The details of the argument leading from Donaldson-
Witten to Seiberg-Witten are at the deepest levels of theoretical physics and,
alas, quite beyond the powers of your poor author whose only service can be
to refer the stout-hearted among his readers to [W3]. The end result, however,
was what Witten was convinced must be a “substitute” for Donaldson theory.
It contained new fields (not a single SU (2)-connection, but a U (1)-connection
and a spinor field), new equations (not the anti-self-dual equations, but the
Seiberg-Witten equations which are, from the point-of-view of partial differ-
ential equations, much simpler), a new moduli space and new invariants. But
the physical equivalence of the quantum field theories from which they arose
(“duality”) left no doubt in Witten’s mind that they must contain the same
topological information. The story of how this conjecture was sprung on the
A.2. Clifford Algebra and Spinc -Structures 363
mathematical community and the pandemonium that ensued has been told
many times, but is best heard from someone who was there so for this as well
as a lovely introduction to what is to come here and a pleasant afternoon’s
entertainment we heartily recommend [Tau2]. We will now get on with the
business of describing the mathematical side of the new classical gauge theory.
When it is possible, with the background we have at our disposal, to provide
details we will do so; when it is not, we will try to provide a sense of what is
involved, what needs to be learned and where it can be learned.
(this is, of course, not a subalgebra of H2×2 ). Notice that det x = ∥q∥2 so,
defining a norm on the set of x given by (A.2.2) by
where we use 1 generically for the identity matrix of any size (2 ×2 in this
case). Note that it follows from (A.2.5) that
xy + yx = −2⟨x, y⟩, x, y ∈ R4 . (A.2.6)
The real subalgebra of H 2×2
generated by {e1 , e2 , e3 , e4 } is the real
Clifford algebra of R4 and is denoted Cl(4). Writing out products of basis
vectors and using (A.2.5) to eliminate linear dependencies gives the following
basis for Cl(4): ( )
1 0
e0 = =1
0 1
( ) ( ) ( ) ( )
0 1 0 i 0 j 0 k
e1 = e2 = e3 = e4 =
−1 0 i 0 j 0 k 0
( ) ( ) ( )
i 0 j 0 k 0
e1 e2 = e1 e3 = e1 e4 =
0 −i 0 −j 0 −k
( ) ( ) ( )
k 0 −j 0 i 0
e2 e3 = e2 e4 = e3 e4 = (A.2.7)
0 k 0 −j 0 i
( ) ( )
0 k 0 −j
e1 e2 e3 = e1 e2 e4 =
−k 0 j 0
( ) ( )
0 i 0 −1
e1 e3 e4 = e2 e3 e4 =
−i 0 −1 0
( )
−1 0
e1 e2 e3 e4 =
0 l
A.2. Clifford Algebra and Spinc -Structures 365
Thus, dim Cl(4) = 16. Since H2×2 itself has real dimension 16 we conclude
that, in fact,
Cl(4) = H2×2 . (A.2.8)
Notice that the basis (A.2.7) gives Cl(4) a natural Z2 -grading
( ) ( ) ( )
q11 q12 q11 0 0 q12
= + .
q21 q22 0 q22 q21 0
for each p ∈ Cl(4). This is clearly an algebra isomorphism that preserves the
grading (A.2.9). Note that if x ∈ R4 ⊆ Cl(4) has ∥x∥ = 1, then, for every
v ∈ R4 ⊆ Cl(4),
adx (v) = xvx−1 = xv(−x) = −xvx
so the identity vx + xv = −2⟨v, x⟩1 implies
In particular,
adx : R4 −→ R4 (x ∈ R4 , ∥x∥ = 1).
But any element of Pin(4) is a product of elements x ∈ R4 with ∥x∥ = 1 so
Spin(u) = adu .
Since any reflection can clearly be written as −adx for some x ∈ R4
with ∥x∥ = 1 and since any rotation can be written as a product of an
even number of reflections, the spinor map is a surjective group homomor-
phism. Finally, to see that ker(Spin) = Z2 = {±1}, so that it is precisely
two-to-one, note that adu is the identity in SO(4) if and only if uxu−1 = x
for each x ∈ R4 . But then u must commute with everything in Cl(4), i.e.,
u ∈ Z(Cl(4)). By Lemma A.2.1, u = ae0 = a1 for some a ∈ R. By (A.2.11),
a2 = 1 and u = ±1.
Remark: Globalizing these constructions leads to the notion of a “spin
structure” on a manifold. In the context of spacetime (as opposed to
Riemannian) manifolds this is just the “spinor structure” we introduced and
studied in Sections 3.5 and 6.5 so we will just briefly describe the defini-
tion and then explain why we need the more general concept of a “spinc
structure”. We let B denote a compact, oriented, smooth 4-manifold with a
πSO
Riemannian metric g . Let SO(4) ,→ FSO (B) −→ B denote the correspond-
ing oriented, orthonormal frame bundle. A spin structure S consists of a
principal Spin(4)-bundle
π
Spin(4) ,→ S(B) −→
S
B
over B and a smooth map
λ : S(B) −→ FSO (B)
satisfying
πSO ◦ λ = πS (A.2.18)
and
λ(p · u) = λ(p) · Spin(u) (A.2.19)
368 Appendix
The fibers of FSO (B) are copies of SO(4) so (A.2.18) says that we have a copy
of Spin(4) “above” each of these and (A.2.19) says that the map λ of S(B)
onto FSO (B) is essentially the spinor map at each point of B. Now, unlike
the frame bundle FSO (B), which exists for any manifold of the type we have
described, there is an obstruction to the existence to a spin structure. The
arguments of Section 6.5 carry over verbatim to show that B admits a spin
structure if and only if the 2nd Stiefel-Whitney class w2 (B) ∈ Ȟ 2 (B; Z2 ) is
trivial. Unfortunately, many interesting 4-manifolds (e.g., CP2 ) do not satisfy
this condition and without a spin structure one cannot define “spinor fields”
in the usual sense. Since spinor fields are crucial to Seiberg-Witten theory and
since one would like this theory to apply to as many 4-manifolds as possible
we seek a generalized notion of both “spin structure” and “spinor field”. As it
happens, there is a very natural generalization obtained by complexifying our
previous algebraic considerations.
To define complex analogues of the algebraic objects we have introduced
we will embed Cl(4) into a complex algebra of matrices and form the complex
subalgebra it generates. The basic tool we use is the usual matrix model of
the quaternions. Specifically, we consider the map γ : H −→ C2×2 from the
quaternions to the 2 × 2 complex matrices given by
( )
α β
γ(q) = γ(α + βj) = (A.2.20)
−β̄ ᾱ
where we have written
q = q 1 + q 2 i + q 3 j + q 4 k = (q 1 + q 2 i) + (q 3 + q 4 i)j = α + βj.
One easily verifies that γ is real linear, injective, preserves products, carries q̄
⊤
to γ(q) and satisfies det(γ(q)) = ∥q∥2 so ( that we) can identify H with the set
α β
of all 2 × 2 complex matrices of the form −β̄ ᾱ . More specifically, if we let
( ) ( )
1 0 i 0
γ(1) = =1 γ(i) = =I
0 1 0 −i
( ) ( ) (A.2.21)
0 1 0 i
γ(j) = =J γ(k) = =K
−1 0 i 0
A.2. Clifford Algebra and Spinc -Structures 369
q = q 1 1 + q 2 I + q 3 J + q 4 K. (A.2.22)
where each γ(qij ) is a 2 × 2 block in the matrix on the right-hand side. This
map Γ is also real linear, injective and preserves products so we can identify
the real algebra Cl(4) with its image.
Cl(4) = Γ(H2×2 )
satisfying
and {( ) }
A1 0
spin(4) = : A1 , A2 ∈ su(2) (A.2.28)
0 A2
corresponding to (A.2.11) and (A.2.12).
Now we regard C4×4 as a complex algebra and define the complexified
Clifford algebra Cl(4) ⊗ C to be the complex subalgebra generated by
{E1 , E2 , E3 , E4 }, i.e., by Cl(4). A basis over C is given by (A.2.7), with all of
the ei capitalized. Since C4×4 also has dimension 16 over C we conclude that
Now let
SC = C4
be the complex vector space C4 with its usual Hermitian inner product
(⟨z, w⟩ = z̄ 1 w1 + z̄ 2 w2 + z̄ 3 w3 + z̄ 4 w4 )) and identify Cl(4) ⊗ C with the
vector space EndC (SC ) of complex linear transformations of SC to itself:
Thus, the elements of Cl(4) ⊗ C (and therefore also Cl(4), R4 and Spin(4))
act as endomorphisms of SC . This action is called Clifford multiplication
and will be written with a dot ·. In particular, we have a representation of the
real Clifford algebra by endomorphisms of SC :
SC ∼ +
= SC −
⊕ SC
z1 z1 0
2 2
z z 0
= + , (A.2.31)
z 3 0 z 3
z4 0 z4
A.2. Clifford Algebra and Spinc -Structures 371
then Clifford multiplication by elements of Cl0 (4), because they are block
+ −
diagonal, preserves SC and SC , whereas Clifford multiplication by elements
+ −
of Cl1 (4), because they are block anti-diagonal, interchanges SC and SC . In
particular, ∆C resolves into a direct sum
−
C ⊕ ∆C
∆C = ∆+
where
∆± ±
C : Spin(4) −→ SU (SC )
−
(see (A.2.27) for the “SU”). ∆+ C and ∆C are inequivalent, irreducible repre-
sentations of Spin(4). Notice also that Clifford multiplication by the elements
of R4 , which are odd, interchanges SC +
and SC−
(this will be crucial when we
define the “Dirac operator” shortly).
Recall that Spin(4) is the set of all even elements in the subgroup of mul-
tiplicative units in the Clifford algebra Cl(4) generated by the unit sphere
in R4 ⊆ Cl(4). For the complex analogue we add to the generators the unit
circle in C. More precisely, we identify U (1) with the subset
U (1) = {eθ i 1 : θ ∈ R}
of Cl(4) ⊗ C (often dropping the “ 1” and thinking of eθ i as an element of
Cl(4) ⊗ C). Then
Spinc (4)
is defined to be the subgroup of the group of multiplicative units in Cl(4) ⊗ C
generated by Spin(4) and U (1). Notice that the elements of Spinc (4) are nec-
essarily even, i.e., in Cl0 (4) ⊗ C. Since U (1) is in the center of Cl(4) ⊗ C we
have
Spinc (4) = {eθ i u : θ ∈ R, u ∈ Spin(4)} (A.2.32)
{ ( ) }
U1 0
= eθ i : θ ∈ R, U1 , U2 ∈ SU (2) .
0 U2
There is yet another useful way of looking at Spinc (4). The mapping
Spinc (4) ∼
= Spin(4) × U (1)/Z2 . (A.2.35)
Finally, notice that, from Lemma A.2.1 and (A.2.32) it follows that the center
of Spinc (4) is
Z (Spinc (4)) = U (1). (A.2.36)
Globalizing all of this to 4-manifolds will require a few mappings which we
now introduce. First define
δ : Spinc (4) −→ U (1)
as follows: For
( ) ( )
U+ 0 eθ i U1 0
ξ= = ∈ Spinc (4),
0 U− 0 eθ i U2
(see (A.2.34)).
One of the Seiberg-Witten equations relates the self-dual part of the curva-
ture of a U (1)-connection to a certain trace free endomorphism of a positive
spinor. The last of our algebraic preliminaries describes the relationship be-
tween 2-forms and endomorphisms. ∧ We note that there is a natural linear
isomorphism from the space 2 (R4 , C) of (complex-valued) 2-forms on R4
into Cl(4) ⊗ C. Indeed, if {e1 , e2 , e3 , e4 } is the standard basis for R4 and
{e1 , e2 , e3 , e4 } is its dual, then we define
∧2
ρ: (R4 , C) −→ Cl0 (4) ⊗ C
by
∑ ∑
ρ(η) = ρ ηij ei ∧ ej = ηij Ei Ej =
i<j i<j
(η12 + η34 )I
+(η13 − η24 )J 0 (A.2.42)
+(η14 + η23 )K
.
(−η12 + η34 )I
0 +(−η13 − η24 )J
+(−η14 + η23 )K
real-valued,
T ∑ T ∑
ρ(η) = η̄ij Ei Ej = ηij ĒjT ĒiT
i<j i<j
∑ ∑
= ηij (−Ej )(−Ei ) = ηij Ej Ei
i<j i<j
∑
= ηij (−Ei Ej )
i<j
= −ρ(η).
Note also that, in (A.2.42), {ea } can be replaced by any oriented, orthonormal
basis provided {Ea } is replaced by its image under Γ (see (A.2.25)).
±
Now, being even (i.e., block diagonal) any ρ(η) preserves the subspaces SC
±
of SC and so we obtain endomorphisms of SC by setting
±
ρ± (η) = ρ(η) | SC . (A.2.43)
+
For example, suppressing the two zero entries in SC (see (A.2.31)),
ρ+ (η) = (η12 + η34 )I + (η13 + η42 )J + (η14 + η23 )K. (A.2.44)
Thus, we have two maps
∧2
ρ± : (R4 , C) −→ EndC (SC
±
). (A.2.45)
∧2 4 ∧2 ∧2
Now let (R , C) = + (R4 , C) ⊕ − (R4 , C) be the decomposition of
∧2 4
(R , C) into self-dual and anti-self-dual 2-forms (relative to the Hodge star
∗ for the usual orientation and inner product on R4 ). We show that ρ± carries
∧2
± (R , C) isomorphically onto the space End0 (SC ) of trace free (complex)
4 ±
±
endomorphisms of SC .
∧2
Lemma A.2.4 ρ± | ± (R4 , C) is a complex linear isomorphism onto
±
End0 (SC ).
∧2 ∧2
Proof: We give the argument for ρ+ | + (R4 , C). The ρ− | − (R4 , C) case
is analogous. A simple computation from (A.2.42) gives
( )
I 0
ρ(e ∧ e + e ∧ e ) = 2
1 2 3 4
0 0
( )
J 0
ρ(e ∧ e + e ∧ e ) = 2
1 3 4 2
(A.2.46)
0 0
( )
K 0
ρ(e ∧ e + e ∧ e ) = 2
1 4 2 3
.
0 0
A.2. Clifford Algebra and Spinc -Structures 375
Since {e1 ∧e2 +e3 ∧e4 , e1 ∧e3 +e4 ∧e2 , e1 ∧e4 +e2 ∧e3 } spans the set of self-
∧2
dual 2-forms on R4 , it is clear that ρ+ | + (R4 , C ) is a linear, injective map
+
to End(SC ). Because I, J and K are trace free, so is everything in the image
∧
of ρ | + (R4 , C). Furthermore, one can show that every 2×2 complex, trace
+ 2
∧2
free matrix is a complex linear combination of I, J and K so ρ+ | + (R4 , C )
maps onto End0 (SC +
).
It ∧ follows, in ∧2 particular, from Lemma A.2.4 that the map
ρ+ | + (R4 , C) : + (R4 , C) −→ End0 (SC
2 +
) has an inverse that we will simply
denote ∧2
+
σ + : End0 (SC ) −→ (R4 , C). (A.2.47)
+
One can compute this inverse explicitly, but we will content ourselves with
describing its action on the particular type of trace free endomorphism that
arises in the Seiberg-Witten equations. For this we consider an element
( )
ψ1
ψ=
ψ2
+ +
of SC (temporarily suppress the two zero components in SC ). Define an en-
+
domorphism of SC by the matrix
( ) ( )
∗ ψl |ψ 1 |2 ψ 1 ψ̄ 2
ψ⊗ψ = 1 2
(ψ̄ ψ̄ ) = . (A.2.48)
ψ2 ψ̄ 1 ψ 2 |ψ 2 |2
satisfying
πSO ◦ Λ = πS c (A.2.53)
and
for each p ∈ S c (B) and each ξ ∈ Spinc (4). Here π : Spinc (4) −→ SO(4) is
defined by (A.2.38).
It is known that, for any B of the type we have described and any choice
of the Riemannian metric g , spinc structures exist (see [LM]). In terms of
transition functions this means that for any trivializing cover {Uα } for the
frame bundle with transition functions gαβ : Uα ∩ Uβ −→ SO(4), there exist
lifts g̃αβ : Uα ∩ Uβ −→ Spinc (4)
A.2. Clifford Algebra and Spinc -Structures 377
Spinc(4)
αβ
π
Uα ∩ Uβ SO (4)
αβ
S(L) = S c (B) ×∆
ˆ C SC
±
S ± (L) = S c (B) ×∆
ˆ ± SC
C
L(L) = S (B) ×δ C
c
S(L) is called the spinor bundle of L, S ± (L) are the positive and
negative spinor bundles of L and L(L) is the determinant line bundle
of L. The algebraic decomposition (A.2.31) persists in the bundle setting to
give a Whitney sum decomposition
associated to L(L).
Remark: This bundle can be described as follows. Choose a Hermitian fiber
metric (smoothly varying Hermitian inner products on the fibers) on the com-
plex line bundle L(L). Then L0 (L) is the unit circle bundle in L(L) i.e., it
is the corresponding oriented, orthonormal frame bundle. One can retrieve
L(L) from L0 (L) as the vector bundle associated to L0 (L) by complex multi-
plication. One can show that w2 (B) = c1 (L0 (L)) mod 2, where w2 (B) is the
second Stiefel-Whitney class of B, and that, conversely, given a U (1)-bundle
L0 over B with w2 (B) = c1 (L0 ) mod 2 there is a spinc structure L on B with
378 Appendix
L0 (L) = L0 . More details on this and much of what follows are available in
[M1]
We will also require a bundle associated to the frame bundle that does not
require a spin or spinc structure. Notice that Spin(4), being contained in
Cl× (4), acts on Cl(4) by conjugation and, since (−u)p(−u)−1 = upu−1 ,
this gives an action of SO(4) = Spin(4)/Z2 on Cl(4) which clearly preserves
products (u(pq)u−1 = (upu−1 )(uqu−1 )). The Clifford bundle Cl(B) is the
bundle with typical fiber Cl(4) over B associated to the frame bundle by this
action.
Cl(B) = FSO (B)×SO(4) Cl(4)
Similarly, one has a complexified Clifford bundle
Cl(B) ⊗ C = FSO (B)×SO(4) (Cl(4) ⊗ C).
These decompose into even and odd summands, e.g., Cl(B) ∼ = Cl0 (B) ⊕
Cl1 (B). Moreover, pointwise multiplication provides the spaces of sections of
these bundles with algebra structures and such sections act on sections of the
spinor bundle by pointwise Clifford multiplication.
Now, Spinc (4) double covers SO(4) × U (1) by the map Spinc so S c (B)
.
double covers the fiber product FSO (B) × L0 (L) (this is just that part of
the product bundle SO(4) × U (1) ,→ FSO (B) × L0 (L) ,→ B × B above the
diagonal in B × B with this diagonal identified with B in the obvious way).
We will use the symbol Spinc also for this double cover.
Spinc (4) ,→ S c (B) −→ B
Spinc (A.2.55)
y
.
SO(4) × U (1) ,→ FSO (B) × L0 (L) −→ B
however.
Recall (Section 3.3) that the frame bundle SO(4) ,→ FSO (B) −→ B has a
distinguished (Levi-Civita) connection which we will denote ω LC . This can
be characterized locally as follows. If {e1 , e2 , e3 , e4 } is a local oriented, or-
thonormal frame field on B (i.e., a section of FSO (B)) with dual 1-form field
{e1 , e2 , e3 , e4 }, then ω LC is represented by a skew-symmetric matrix (ω ij )
of R-valued 1-forms satisfying dei = −ω ij ∧ ej , i = 1, 2, 3, 4. Now no-
tice that if B had a spin structure Spin(4) ,→ S(B) −→ B, then the map
λ : S(B) −→ FSO (B) is a double cover that respects the group actions so
that any connection on FSO (B), e.g., ω LC , automatically lifts to a connec-
tion on S(B) (think of the connection as a distribution of horizontal spaces).
However, if B has only a spinc structure Spinc (4) ,→ S c (B) −→ B, then the
map Λ of (A.2.52) is not a finite covering so ω LC alone will not determine a
connection on S c (B). However, Spinc : S c (B) −→ FSO (B)×L ˙ 0 (L) is a dou-
0
ble cover (A.2.55) and if A is any connection on L (L), then A and ω LC
together determine a connection on FSO (B)×L ˙ 0 (L) which will then lift to
c
a connection on S (B). Specifically, if prF and prL0 denote the restrictions
to FSO (B) × ˙ L0 (L) of the projections of FSO (B) × L0 (L) onto FSO (B) and
0
L (L), respectively, then
prF ∗ ω LC ⊕ prL0 ∗A
is a connection on the fiber product and, identifying spinc (4) with the subset
of Cl(4) ⊗ C given in (A.2.40),
ω A = (Spinc )∗ (prF ∗ ω LC ⊕ prL0 ∗A) (A.3.1)
380 Appendix
̸ ˜ A : Γ(S(L)) −→ Γ(S(L))
D
and
( ) ( )
̸ A∗ : Γ S − (L) −→ Γ S + (L)
D (A.3.4)
(these are, in fact, adjoints relative to the L2 inner product on sections induced
by the pointwise Hermitian inner product on fibers). We will also follow the
custom in mathematics of referring to D ̸ A also as a Dirac operator.
With this we can (at last) formulate the Seiberg-Witten equations. Thus,
we let B denote a compact, connected, simply connected, oriented, smooth
4-manifold. Select a Riemannian metric g for B and then a spinc structure L
for the corresponding oriented, orthonormal frame bundle FSO (B). A pair
(A, ψ) consisting of a U (1)-connection A on U (1) ,→ L0 (B) −→ B and
a positive spinor field ψ ∈ Γ(S + (L)) satisfies the Seiberg-Witten (SW)
equations if
̸ Aψ = 0
D (Dirac Equation) (A.3.5)
and
∗
A = σ ((ψ ⊗ ψ )0 )
F+ +
(Curvature Equation), (A.3.6)
where F +
A is the g -self-dual part of the curvature of A.
4
Λ 4
× Spinc(4) × SO(4)
4
382 Appendix
where Λ(b, ξ) = (b, π(ξ)) with π given by (A.2.38). The spinor bundles are
therefore also trivial so their sections can be identified with globally defined
functions on R4 which we will write
ψ1
2
ψ
Ψ = 3 : R4 −→ SC ∼ = C4
ψ
ψ4
ψ1
2
ψ + ∼
ψ= : R4 −→ SC = C2
0
0
0
0 − ∼
ϕ = 3 : R4 −→ SC = C2 .
ψ
ψ4
( )
ψ
For convenience we will often abuse the notation and write Ψ = ϕ by
1 2 3 4
suppressing the zero components. We use x , x , x , x for the standard coor-
dinates on R4 and write ∂i for ∂x ∂
i , i = 1, 2, 3, 4 (these being applied compo-
Thus, with {ei } = {∂i } the standard oriented, orthonormal frame field on R4 ,
we have ∇A Ψ(ei ) = ∂i Ψ + Ai Ψ and, for convenience, we will write this as
∂i ψ 1 + Ai ψ 1
∂i ψ 2 + Ai ψ 2
∇i Ψ = (∂i + Ai )Ψ = ∂ ψ3 + A ψ3 .
i i
∂i ψ 4 + Ai ψ 4
∑4
The Dirac operator D ̸ ˜ A Ψ = i=1 ei · ∇i Ψ requires that we Clifford multiply
by the basis elements ei , i.e., matrix
( multiply
) by Ei = Γ(ei ) ∈ Cl(4) ⊗ C as
in (A.2.25). For this we write Ψ = ψϕ so that
∑
4 ∑
4
̸ ˜ AΨ =
D ei · ∇i Ψ = Ei ∇i Ψ
i=1 i=1
( )
∇1 ϕ + I∇2 ϕ + J∇3 ϕ + K∇4 ϕ
= .
−∇1 ψ + I∇2 ψ + J∇3 ψ + K∇4 ψ
( )
Note that, as expected, D ̸ ˜ A sends positive spinor fields Ψ = ψ0 to negative
̸ ˜ A to positive spinor fields will
spinor fields and vice versa. The restriction of D
be written
̸ A ψ = −∇1 ψ + I∇2 ψ + J∇3 ψ + K∇4 ψ
D (A.3.7)
by suppressing the zero components (but now one must remember that D ̸ Aψ
is a negative spinor field). The first Seiberg-Witten equation (A.3.5) then
becomes
∇1 ψ = I∇2 ψ + J∇3 ψ + K∇4 ψ (A.3.8)
1 1
F+
A = (F A + ∗ F A ) = (F12 + F34 )(dx1 ∧ dx2 + dx3 ∧ dx4 )
2 2
1
+ (F13 + F42 )(dx1 ∧ dx3 + dx4 ∧ dx2 )
2
1
+ (F14 + F23 )(dx1 ∧ dx4 + dx2 ∧ dx3 ).
2
Thus, (A.3.6) becomes
1
F12 + F34 = − ψ ∗ Iψ
2
1
F13 + F42 = − ψ ∗ Jψ (A.3.10)
2
1
F14 + F23 = − ψ ∗ Kψ
2
or, in more detail,
1 ( )
(∂1 A2 − ∂2 A1 ) + (∂3 A4 − ∂4 A3 ) = − i |ψ 1 |2 − |ψ 2 |2
2 ( )
(∂1 A3 − ∂3 A1 ) + (∂4 A2 − ∂2 A4 ) = −i Re ψ̄ 1 ψ 2 (A.3.11)
( )
(∂1 A4 − ∂4 A1 ) + (∂2 A3 − ∂3 A2 ) = −i Im ψ̄ 1 ψ 2 .
d(F1 (dx1 ∧ dx2 − dx3 ∧ dx4 )) = dF1 ∧ (dx1 ∧ dx2 − dx3 ∧ dx4 )
= (∂1 F1 dx1 + ∂2 F1 dx2 + ∂3 F1 dx3 ) ∧ (dx1 ∧ dx2 − dx3 ∧ dx4 )
= −(∂1 F1 )dx1 ∧ dx3 ∧ dx4 − (∂2 F1 )dx2 ∧ dx3 ∧ dx4 +(∂3 F1 )dx1 ∧ dx2 ∧ dx3 .
Computing the remaining terms similarly one finds that the coefficient of dx1 ∧
dx3 ∧ dx4 is −(∂1 F1 + ∂2 F2 + ∂3 F3 ) and this is just (minus) the divergence of
the map F = (F1 , F2 , F3 ). The remaining coefficients are just the components
A.4. The Moduli Space and Invariant 385
and
Spinc (4) ,→ S c (B) −→ B
prL0 ◦ Spinc (A.4.2)
y
U (1) ,→ L0 (L) −→ B
The collection of all such is a group G(L) under composition which we call
the (Seiberg-Witten) gauge group and which we will show acts naturally
on the solutions to (SW).
Lemma A.4.1 If γ ∈ C ∞ (B, U (1)) is any smooth map of B into
U (1) ⊆ Spinc (4), then the map
σγ : S c (B) −→ S c (B)
σγ (p) = p · γ(πS c (p))
G(L) ∼
= C ∞ (B, U (1))
i.e.,
ψ · σγ = ψ · γ = σγ ∗ ψ = ψ ◦ σγ . (A.4.4)
The same formulas define the action of G(L) on negative spinor fields.
Turning next to the connection A on U (1) ,→ L0 (L) −→ B we note that the
automorphism σγ of S c (B) induces an automorphism σ ′ γ of L0 (L) as follows:
σγ
S c (B) S c (B)
L0( ) L0( )
σ ɂγ
(we will write out an explicit local expression for σ ′ γ shortly). Now we define
the action of σγ ∈ G(L) on A by
A · σγ = A · γ = (σ ′ γ )∗ A. (A.4.8)
it satisfies
( ) (( ) )
prL0 ◦ Spinc (p · ξ0 ) = prL0 ◦ Spinc (p) · δ(ξ0 ) (A.4.10)
so
(( ) ) ( )( )
σ′ γ prL0 ◦ Spinc (p) = prL0 ◦ Spinc σγ (p)
( )( )
= prL0 ◦ Spinc p · γ(πS c (p))
(( ) ) ( )
= prL0 ◦ Spinc (p) · δ γ(πS c (p)) .
388 Appendix
and therefore
A · γ = A + 2γ −1 dγ. (A.4.13)
or, locally on B,
In order to show that this action preserves solutions to (SW) we first observe
that the spinc connection corresponding to A · γ is the pullback by σγ of that
corresponding to A, i.e.,
ω A·γ = σγ ∗ ω A . (A.4.16)
and
( ( ))
ω A·γ = (Spinc )∗ prF∗ ω LC + prL∗0 (σγ ′ )∗ A
= (prF ◦ Spinc )∗ ω LC + (σγ ′ ◦ prL0 ◦ Spinc )∗ A
= (prF ◦ Spinc ◦ σγ )∗ ω LC + (prL0 ◦ Spinc ◦ σγ )∗ A
A.4. The Moduli Space and Invariant 389
Proof: For the curvature equation we observe that (A.4.16), the usual trans-
formation equation for the curvature and the fact that U (1) is Abelian imply
that
+
F A·γ = γ 2 F A+ (γ 2 )−1 = F A+ .
Similarly, the commutativity of U (1) gives
(ψ · γ) ⊗ (ψ · γ)∗ = (γ −1 ψ) ⊗ (γ −1 ψ)∗
= (γ −1 ψ) ⊗ (γ ψ ∗ )
= (γ −1 γ)(ψ ⊗ ψ ∗ )
= ψ ⊗ ψ∗ .
+
For this it will be convenient to identify ψ with an equivariant SC -valued map
on S (B) and compare the covariant exterior derivatives dA ψ and dA·γ (ψ · γ).
c
There are standard formulas for such derivatives (see, for example, (6.8.4) of
[N4]) which, in our case, give
1
dA ψ = dψ + ω A ψ
2
and
1
dA·γ (ψ · γ) = d(ψ · γ) + ω A·γ (ψ · γ) ,
2
where, e.g., ω A takes values in spin (4), identified with a subset of Cl(4) ⊗ C
c
Finally,
∑
4
̸ A·γ (ψ · γ) =
D ei · ∇A·γ (ψ · γ)(ei )
i=1
∑4
= ei · (ψ ◦ πS c )−1 ∇A ψ (ei )
i=1
∑
4
−1
= (ψ ◦ π ) Sc ei · ∇A ψ (ei )
i=1
= (ψ ◦ πS c )−1 D
̸ Aψ
= (̸DA ψ) · γ
(A + 2 γ −1 d γ, γ −1 ψ) = (A, ψ)
γ −1 ψ = ψ and 2 γ −1 d γ = 0 .
Since γ ̸= 1, the first of these can be true if and only if ψ ≡ 0. The second
implies d γ = 0 and, since B is connected, γ must be constant.
An (A, ψ) ∈ A(L) is said to be reducible if ψ ≡ 0 and irreducible
otherwise.
As was the case in Donaldson theory, the configuration space A(L) is an
affine space and therefore an infinite-dimensional manifold. The tangent space
at any (A, ψ) ∈ A(L) can be identified with
∧1
T(A,ψ) (A(L)) = (B, Im C) ⊕ Γ(S + (L)) (A.4.21)
(no adjoint bundle required in the first summand because U (1) is Abelian).
The gauge group G(L) has the structure of a Hilbert Lie group whose Lie
∧0
algebra can be identified with (B, Im C). Fixing (A, ψ) ∈ A(L), the ac-
tion of G(L) on (A, ψ) gives a map G(L) −→ A(L) whose derivative at the
392 Appendix
identity is
∧0 ∧1
(B, Im C) −→ (B, Im C) ⊕ Γ(S + (L))
a −→ (2 d a, −a · ψ) , (A.4.22)
d( ) d( )
(A, ψ) · ei tθ(x) = A + 2i tdθ, e−i tθ(x) ψ
dt t=0 dt t=0
= (2i dθ, −iθ(x)ψ)
= (2d a, −aψ)
= (2d a, −a · ψ) .
B(L) = A(L)/G(L)
of configurations is a smooth Banach manifold away from the reducible config-
urations (i.e., away from those [A, ψ] with ψ ≡ 0) and the monopole moduli
space M(L) is a subset of it. Define the Seiberg-Witten map
∧2
F : A(L) −→ (B, Im C) ⊕ Γ(S − (L))
+
by
F (A, ψ) = (F A+ − σ + ((ψ ⊗ ψ ∗ )0 ), D
̸ A ψ) . (A.4.23)
Then (A, ψ) satisfies (SW ) if and only if F (A, ψ) = (0, 0). The derivative of
F at (A, ψ) ∈ M(L), i.e., the linearization of the Seiberg-Witten equations
at (A, ψ), is
∧1 ∧2
F∗(A,ψ) : (B, Im C) ⊕ Γ(S + (L)) −→ (B, Im C) ⊕ Γ(S − (L))
+
( )
d+ −Dψ
F∗(A,ψ) = (A.4.24)
· 12 ψ ̸ A
D
where
∧1 ∧2
d+ : (B, Im C) −→ (B, Im C)
+
A.4. The Moduli Space and Invariant 393
(the object inside the parentheses being a section of End0 (S + (L)) which σ +
identifies with a self-dual 2-form on B).
Remark: (A.4.24) can be verified with a local argument analogous to that
for (A.4.22).
Associated to any solution (A, ψ) to (SW ) is a fundamental elliptic
complex E(A, ψ):
∧0 ∧1
0 −→ (B, Im C) −→ (B, Im C) ⊕ Γ(S + (L))
∧2
−→ (B, Im C) ⊕ Γ(S − (L)) −→ 0 ,
+
where the second and third maps are, respectively, the derivative (A.4.22) of
the action of G(L) on (A, ψ), and the derivative (A.4.24) of the Seiberg-
Witten map at (A, ψ). This complex has finite-dimensional cohomology
groups H i (A, ψ), i = 0, 1, 2, which admit interpretations analogous to those
in Donaldson theory:
H 0 (A, ψ) = tangent space to the stablizer of (A, ψ) in G(L) so
H 0 (A, ψ) = 0 ⇐⇒ (A, ψ) irreducible
⇐⇒ ψ ̸≡ 0
Exactly as in the case of Donaldson theory one can show that, for a fixed
generic metric g and perturbation η and any associated spinc structure L,
a choice of orientation for the vector space H+ 2
(B; R ) canonically orients
all of the moduli spaces M(L, η). Likewise as in Donaldson theory, when
b+
2 (B) > 1 there is a cobordism result which roughly says that for a generic
l-parameter family g (t), 0 ≤ t ≤ 1, of metrics and a generic l-parameter
family η(t), 0 ≤ t ≤ 1, of perturbations, the moduli spaces parametrized
by t fit together to form a smooth manifold with boundary containing no
points corresponding to reducible solutions. The boundary is the disjoint union
of moduli spaces for (g (0), η(0)) and (g (1), η(1)). Moreover, selecting an
orientation for H+2
(B; R) orients this parametrized moduli space and the two
boundary moduli spaces inherit opposite orientations.
Remark: There is a technical point which we glossed over here and should
mention because it has no analogue in Donaldson theory. Changing the metric
g changes the orthonormal frame bundle and so, one would think, the spinc
structure. It would appear that the discussion above is incomplete without a
specification of a spinc structure for each t. In fact, however, one can show
that frame bundles for different metrics are naturally isomorphic and so one
can pull back spinc structures by the isomorphisms, thus effectively “fixing”
L (up to equivalence) regardless of the choice of g .
Except for a few minor simplifications and adaptations the story of the
Seiberg-Witten moduli space thus far has been virtually indistinguishable
from what we had to say about the anti-self-dual moduli space. The one
aspect of Seiberg-Witten theory that differs significantly from Donaldson the-
ory (and that accounts for its relative simplicity) is that there is no need for
an “Uhlenbeck- style compactification”:
For any metric g , and spinc structure L
and any perturbation η, the moduli space
M(L, η) is always compact.
396 Appendix
(note that (A.4.31) implies that ⟨ ∇A∗ ◦ ∇A ψ, ψ ⟩ must be real). The proof
depends on the identity
k(B) = max{k(x0 ) : x0 ∈ B}
and conclude that for any fixed metric and any spinc structure, any solution
(A, ψ) to the Seiberg-Witten equations has spinor field ψ bounded by the
geometrical constant k(B):
(such an L need not exist). Assuming that an orientation for the vector space
H+2
(B; R) has been fixed, the moduli space is a finite set of isolated points each
of which is equipped with a sign ±1. The sum of these signs is an integer and,
as in Donaldson theory, when b+ 2 (B) > 1 a cobordism argument shows that
the integer is independent of the choice of (generic) metric and perturbation.
We call this integer the 0-dimensional Seiberg-Witten invariant of B
associated with L and denote it
invariants. The conjecture was remarkable. It was a very deep and purely
mathematical statement, but one suggested entirely by physics. We will con-
clude with a very brief description of what Witten believed must be true (and
was eventually proved by Feehan and Leness [FeLe]).
The sequence γd (B), d = 0, 1, 2, . . ., of Donaldson invariants for B can be
assembled into a single formal power series
∞
∑ γd (B)(x)
DB (x) =
d!
d=0
on H 2 (B) called the Donaldson series of B. For example, all of the invariants
for K3 have been computed and one finds that
∑∞
1
DK3 (x) = (QK3 (x, x)/2)n = exp(QK3 (x, x)/2).
n=0
n!
for every x ∈ H2 (B). Moreover, each KM-basic class reduces mod 2 to the
second Stiefel-Whitney class w2 (B) of B.
The appearance of Theorem A.5.1 in the spring of 1994 was an extraor-
dinary breakthrough in Donaldson theory. Enormously complicated calcula-
tions of an infinite set of invariants were suddenly replaced by the (certainly
400 Appendix
not easy, but at least finite) problem of determining the KM-basic classes
and coefficients. As fate would have it, however, the fall of 1994 witnessed
another event which rendered this triumph of Kronheimer and
Mrowka moot. Edward Witten, in his now famous lecture at M.I.T.
(described in [Tau2]), made the conjecture which, within weeks, brought
about the demise of Donaldson theory and initiated an entirely new approach
to the study of smooth 4-manifolds.
401
402 References
What follows is a list of the those symbols that are used consistently through-
out the text, a brief discription of their meaning and/or a reference to the
page on which such a description can be found.
R, real numbers
C, complex numbers
H, quaternions
F = R, C, or H
⟨ , ⟩, inner product on Fn
| |, absolute value in R, modulus in C or H
dim X, dimension of the manifold X
C ∞ (X), real-valued C ∞ -functions on X
id, identity map
Tp (X), tangent space at p
Di , i th partial derivative operator
ι, inclusion map
α′ (t0 ), velocity vector to α at t0
f∗p , derivative of f at p
V ∗ , dual of vector space V
(US , φS ), (UN , φN ), stereographic projection charts
∥ ∥, norm in Euclidean space
SU (2), 5
FPn−1 , projective spaces
[ ξ ], equivalence class containing ξ
P, projection map
GL(n, F), 5
GL(n, F), 5
U (n, F), 5
O(n), 5
U (n), 5
Sp(n), 5
SO(n), 5
SU (n), 5
ĀT , conjugate transpose of A
φ × ψ, 5
407
408 Symbols
V (f ) = V f , 7
X (X), smooth vector fields on X
[V , W ], Lie bracket of vector fields
[e1 , . . . , en ], orientation class of {e1 , . . . , en }
Tp∗ (X), cotangent space at p
df , exterior derivative of f
dfp = df (p)
X ∗ (X), smooth 1-forms on X
Θ(V ) = ΘV , 7
F ∗ , pullback by F
⊗, tensor product
∧, wedge product
∧ρ , ρ-wedge product
SL(n, F), 12
[ , ], Lie bracket
trace (A), trace of the matrix A
Im C, pure imaginary complex numbers
Im H, pure imaginary quaternions
G, Lie algebra of the Lie group G
e, identity element in the group G
o(n), 14
so(n), 14
u(n), 14
su(n), 14
sp(n), 14
σ1 , σ2 , σ3 , Pauli spin matrices, 15
exp (A) = eA , 18
σ, right action
p · g, right action of g on p
σg , right action by g
ρ, left action, 21
g · p, left action of g on p
ρg , left action by g
P
G ,→ P −→ X, principal G-bundle over X
Vertp (P ), vertical vectors at p in P
A# , fundamental vector field
σp , left action on p
A, gauge potential
Symbols 409
C ∗ , cochain complex
H k (C ∗ ), k th cohomology group of C ∗
Λ∗ (X), cochain of forms on X
QX , intersection form on X
deg(f ), degree of f
H(f ), Hopf invariant of f
c1 (P ), 1st Chern class of P
S k (V), complex-valued symmetric k-multilinear maps on V
S(V), direct sum of the S k (V)
P k (V), homogeneous polynomials of degree k on V
P (V), direct sum of the P k (V)
Sρk (V), ρ-invariant subspace of S k (V)
Pρk (V), ρ-invariant subspace of P k (V)
Sρ (V), direct sum of the Sρk (V)
Pρ (V), direct sum of the Pρk (V)
I k (G) = Sad
k
(G), 309
I(G), direct sum of the I k (G)
symtr, symmetrized trace
σ
P , 310
S0 , S1 , . . . , Sn , elementary symmetric polynomials
ck (P ), k th Chern class of P
| σ |, support of the simplex σ
N (U), nerve of U
Č j (U; Z2 ), Čech j-cochain group
δ j , coboundary operator
B̌ j (U; Z2 ), Čech j-coboundaries of U
Ž j (U; Z2 ), Čech j-cocycles of U
Ȟ j (U; Z2 ), j th Čech cohomology group of U
Ȟ j (X; Z2 ), j th Čech cohomology group of X
w1 (X), 1st Stiefel-Whitney class of X
w2 (X), 2nd Stiefel-Whitney class of X
γd (M), Donaldson invariants of M
 ×Ĝ Λ2+ (M , ad P ), vector bundle of Donaldson theory, 356
e(X), Euler class of the manifold X
Pf, Pfaffian, 357
e(E), Euler class of the vector bundle E
χ(E), Euler number of the vector bundle E
TQFT, topological quantum field theory
412 Symbols
A C
action canonical isomorphism, 3
effective, 22 Cartan 1-form, 15
free, 22 Cartan Structure Equation, 35
left, 21 Čech cohomology group, 339, 343
right, 21 Čech j-coboundary, 339
transitive, 22 Čech j-cocycle, 339
action functional, 47 characteristic class, 303, 325
Donaldson-Witten, 360 chart, 1
gauge theory, 47 admissible, 1
spin 0-electrodynamics, 60 consistent with orientation, 7
SU (2) Yang-Mills-Higgs, 111 chiral representation, 82
Yang-Mills, 38, 48, 116 Chern-Weil homomorphism, 325
ad(G)-invariant, 20, 21 Chern number, 328
adjoint bundle, 40 Chern class, 325
total, 334
adjoint representation, 19
C ∞ -manifold, 1
admissible basis, 164
C ∞ -map, 2
algebraic homotopy, 219
C ∞ -related, 1
almost everywhere, 238
classical groups, 5
antiparticle, 64
Classification Theorem, 28
antipodal map, 296 Clifford algebra of R4 , 364
anti-self-dual (ASD), 115 center of, 365
Atiyah-Singer Index Theorem, 394 complexified, 370
atlas, 1 even elements of, 365
maximal, 1 odd elements of, 365
oriented, 7 real, 364
automorphism of bundles, 28 Clifford bundle, 378
complexified, 378
Clifford multiplication, 370
B closed form, 217
Betti numbers coboundary, 275
Bianchi Identity, 231 coboundary operator, 275
Bogomolny monopole cochain complex, 275
equations, 114 cochain homotopy, 277
boosts, 167 cochain map, 276
Borel sets, 235 cocycle, 275
BPST potential, 34, 115 cocycle condition, 26
center, 34, 115 cohomologous, 276
instanton, 115 cohomology class, 276
scale, 34, 115 commutator, 13
bump function, 141 complex line bundle, 39
bundle map, 27 complex scalar field, 42
bundle splicing, 99 conjugation representation, 88
413
414 Index
connection, 29 diffeomorphic, 1
existence of, 144 diffeomorphism, 1
Levi-Civita, 160 differentiable manifold, 1
linear, 160 differentiable structure, 1
metric, 161 differential form, 206
Riemannian, 160 vector valued, 211
coordinate expression, 1 dimension
coordinates on R1,3 , 173 of a differentiable manifold, 1
null, 173 of a submanifold, 2
spherical, 172 of a of a topological manifold, 1
cotangent bundle, 154 dimensional reduction, 104
cotangent space, 7 Dirac electron, 102
Coulomb field, 55 coupled, 102
coupling constant,47, 359 free, 99
covariant derivative, 64 Dirac equation, 83
covariant exterior derivative, 225 Dirac magnetic monopole, 37, 104
covariant tensors, 180 Dirac matrices, 82
rank, 180 Dirac quantization condition, 57
rank one, 8 Dirac spinor, 99
rank two, 8 direct sum of cochains, 280
covariant tensor field of rank domain with smooth boundary, 250
two, 9 boundary, 250
components, 9 exterior, 250
continuous, 9 interior, 250
smooth, 9 Dominated Convergence
covector, 7 Theorem, 239
critical value, 2 Donaldson invariants, 352
cross-section, 26 Donaldson µ-map, 354 355
global, 28 Donaldson series, 412
curvature, 34 duality, 293, 362
curvature equation, 381 Duistermaat-Heckman
theorem, 361
D
deformed, 255 E
degree, 294 Einstein cylinder, 177
de Rham, 258 Einstein-de Sitter spacetime, 173
coboundary, 258 electric charge, 53, 55, 56
cocycles, 258 electric field, 51, 55
cohomology group, 258 electromagnetic theory, 52
complex, 216 elementary symmetric
derivation, 6 polynomials, 310
derivative, 2 elevator experiment, 168
de Sitter spacetime, 175 elliptic complex, 393
determinant line bundle, 377 equation of structure, 35
Index 415
M
K
magnetic charge, 37, 334
k-dimensional submanifold, 2 magnetic field, 52
k-tensor magnetic monopole, see 104
skew-symmetric, 180 Dirac magnetic monopole
symmetric, 180 map induced in cohomology, 262
vector-valued, matrix Lie group, 12
Killing form, 20 matter field, 42
Klein-Gordon equation, 63 Maurer-Cartan equations, 16
KM-basic classes, 399 Maxwell’s equations, 53
KM-coefficients, 399 Mayer-Vietoris sequence, 281
KM-simple type, 399 measure zero, 235
measurable form, 243
L measurable function, 237
Laplacian, 216 measurable set, 236
Lebesgue integrable, 238 metric, 20
form, 243 bi-invariant, 20
Lebesgue integral, 237 left invariant, 20
Leaning Tower of Pisa, 168 Lorentz, 169
Leibnitz Product Rule, 2 Riemannian, 9
lens space, 60 right invariant, 20
Lie algebra, 12 semi-Riemannian, 9
ideal in, 20 metric volume form, 190, 209
of a Lie group, 14 Minkowski inner product, 164
Lie bracket, 6 Minkowski space, 155
trivial, 13 Minkowski spacetime, 155
Lie group, 11 moduli space, 351
matrix, 12 Donaldson theory, 352
semisimple, 20 Seiberg-Witten theory, 391
light cone, 164 monopole number, 130
lightlike, 164 Monotone Convergence
local field strength, 35, 46 Theorem, 239
Index 417
V
Y
vacuum state, 113
Yang-Mills action, 38, 48, 115
vector analysis, 216
Yang-Mills equations, 48, 115
vector bundle, 39
Yang-Mills theory, 48, 49, 115
vector field, 5
Yang-Mills-Higgs action, 111
components of, 6
gauge invariance, 112
continuous, 6
Yang-Mills-Higgs equations, 114
horizontal lifts, 228
Yang-Mills-Higgs monopoles, 112
left-invariant, 13
smooth, 6
vector-valued forms, 10, 198 Z
velocity vector, 2 zero section, 272, 356