intro_manifolds (2)
intro_manifolds (2)
Zh Felix Zhou
January 3, 2024
1
x
eli
©F
2
eli
x
Zh
ou
Contents
ou
1 Euclidean Spaces 11
1.1 Smooth Functions on a Euclidean Space . . . . . . . . . . . . . . . . . . . . 11
Zh
1.1.1 Smooth Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.1.2 Taylor’s Theorem with Remainder . . . . . . . . . . . . . . . . . . . 12
1.2 Tangent Vectors in Rn as Derivations . . . . . . . . . . . . . . . . . . . . . . 13
1.2.1 The Directional Derivative . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2.2 Germs of Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2.3 Derivations at a Point . . . . . . . . . . . . . . . . . . . . . . . . . . 15
x
1.2.4 Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.2.5 Vector Fields as Derivations . . . . . . . . . . . . . . . . . . . . . . . 19
1.3 The Exterior Algebra of Multicovectors . . . . . . . . . . . . . . . . . . . . . 20
eli
1.3.1 Dual Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.3.2 Permutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.3.3 Multilinear Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.3.4 The Permutation Action on Multilinear Functions . . . . . . . . . . . 22
©F
3
1.3.10 A Basis for k-Covectors . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.3.11 Useful Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1.4 Differential Forms on Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.4.1 Differential 1-Forms and the Differential of a Function . . . . . . . . 33
1.4.2 Differential k-Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
ou
1.4.3 Differential Forms as Multilinear Functions on Vector Fields . . . . . 37
1.4.4 The Exterior Derivative . . . . . . . . . . . . . . . . . . . . . . . . . 38
1.4.5 Closed & Exact Forms . . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.4.6 Applications to Vector Calculus . . . . . . . . . . . . . . . . . . . . . 41
Zh
1.4.7 Convention on Subscripts and Superscripts . . . . . . . . . . . . . . . 44
1.4.8 Miscellaneous Results . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2 Manifolds 47
2.1 Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.1.1 Topological Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.1.2 Compatible Charts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
x
2.1.3 Smooth Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.1.4 Examples of Smooth Manifolds . . . . . . . . . . . . . . . . . . . . . 51
2.2 Smooth Maps on a Manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
eli
2.2.1 Smooth Functions on a Manifold . . . . . . . . . . . . . . . . . . . . 54
2.2.2 Smooth Maps between Manifolds . . . . . . . . . . . . . . . . . . . . 55
2.2.3 Diffeomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
©F
4
2.3.2 Continuity of a Map on a Quotient . . . . . . . . . . . . . . . . . . . 64
2.3.3 Identification of Subset to a Point . . . . . . . . . . . . . . . . . . . . 64
2.3.4 A Necessary Condition for a Hausdorff Quotient . . . . . . . . . . . . 65
2.3.5 Open Equivalence Relations . . . . . . . . . . . . . . . . . . . . . . . 65
2.3.6 The Real Projective Space . . . . . . . . . . . . . . . . . . . . . . . . 67
ou
2.3.7 The Standard Smooth Atlas on a Real Projective Space . . . . . . . 69
2.3.8 The Grassmannian Manifold . . . . . . . . . . . . . . . . . . . . . . . 70
Zh
3.1.1 The Tangent Space at a Point . . . . . . . . . . . . . . . . . . . . . . 75
3.1.2 The Differential of a Map . . . . . . . . . . . . . . . . . . . . . . . . 76
3.1.3 The Chain Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.1.4 Bases of Tangent Space at a Point . . . . . . . . . . . . . . . . . . . 79
3.1.5 A Local Expression for the Differential . . . . . . . . . . . . . . . . . 80
3.1.6 Curves in a Manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
x
3.1.7 Computing the Differential Using Curves . . . . . . . . . . . . . . . . 84
3.1.8 Immersions and Submersions . . . . . . . . . . . . . . . . . . . . . . 85
3.1.9 Rank, Critical & Regular Points . . . . . . . . . . . . . . . . . . . . . 86
eli
3.1.10 Useful Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.2 Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.2.1 Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
©F
5
3.3.1 Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.3.2 Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
u
3.3.3 The Dual and Multicovector Functors . . . . . . . . . . . . . . . . . 105
3.4 The Rank of a Smooth Map . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3.4.1 Constant Rank Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 107
o
3.4.2 The Immersion & Submersion Theorems . . . . . . . . . . . . . . . . 110
3.4.3 Images of Smooth Maps . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.4.4 Smooth Maps into Submanifold . . . . . . . . . . . . . . . . . . . . . 114
Zh
3.4.5 The Tangent Plane to a Surface in R3 . . . . . . . . . . . . . . . . . 115
3.4.6 The Differential of an Inclusion Map . . . . . . . . . . . . . . . . . . 117
3.4.7 Useful Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
3.5 The Tangent Bundle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3.5.1 The Topology of the Tangent Bundle . . . . . . . . . . . . . . . . . . 119
3.5.2 The Manifold Structure on the Tangent Bundle . . . . . . . . . . . . 122
3.5.3 Vector Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
3.5.4 Smooth Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
ix
3.5.5 Smooth Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.6 Bump Functions and Partitions of Unity . . . . . . . . . . . . . . . . . . . . 129
3.6.1 Smooth Bump Functions . . . . . . . . . . . . . . . . . . . . . . . . . 130
3.6.2 Partitions of Unity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
3.6.3 Existence of a Partition of Unity . . . . . . . . . . . . . . . . . . . . 132
el
6
4 Lie Groups and Lie Algebras 147
4.1 Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.1.1 Lie Groups & Examples . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.1.2 Lie Subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.1.3 The Matrix Exponential . . . . . . . . . . . . . . . . . . . . . . . . . 151
ou
4.1.4 The Trace of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 153
4.1.5 The Differential of Det at the Identity . . . . . . . . . . . . . . . . . 154
4.2 Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.2.1 Tangent Space at the Identity of a Lie Group . . . . . . . . . . . . . 155
Zh
4.2.2 Left-Invariant Vector Fields on a Lie Group . . . . . . . . . . . . . . 157
4.2.3 The Lie Algebra of a Lie Group . . . . . . . . . . . . . . . . . . . . . 159
4.2.4 The Lie Bracket on gl(n, R) . . . . . . . . . . . . . . . . . . . . . . . 160
4.2.5 The Pushforward of Left-Invariant Vector Fields . . . . . . . . . . . . 162
4.2.6 The Differential as a Lie Algebra Homomorphism . . . . . . . . . . . 163
7
5.2.5 Pullback of k-Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
5.2.6 The Wedge Product . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5.2.7 Differential Forms on a Circle . . . . . . . . . . . . . . . . . . . . . . 184
5.2.8 Invariant Forms on a Lie Group . . . . . . . . . . . . . . . . . . . . . 184
5.3 The Exterior Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
ou
5.3.1 Exterior Derivative on a Coordinate Chart . . . . . . . . . . . . . . . 187
5.3.2 Local Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
5.3.3 Existence of an Exterior Derivative on a Manifold . . . . . . . . . . . 189
5.3.4 Uniqueness of the Exterior Derivative . . . . . . . . . . . . . . . . . . 189
Zh
5.3.5 Exterior Differentiation Under a Pullback . . . . . . . . . . . . . . . 191
5.3.6 Restriction of k-Forms to a Submanifold . . . . . . . . . . . . . . . . 193
5.3.7 A Nowhere-Vanishing 1-Form on the Circle . . . . . . . . . . . . . . 194
5.4 The Lie Derivative & Interior Multiplication . . . . . . . . . . . . . . . . . . 195
5.4.1 Families of Vector Fields and Differential Forms . . . . . . . . . . . . 195
5.4.2 The Lie Derivative of a Vector Field . . . . . . . . . . . . . . . . . . 198
5.4.3 The Lie Derivative of a Differential Form . . . . . . . . . . . . . . . . 201
x
5.4.4 Interior Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . 202
5.4.5 Properties of the Lie Derivative . . . . . . . . . . . . . . . . . . . . . 205
eli
5.4.6 Global Formulas for the Lie and Exterior Derivatives . . . . . . . . . 208
6 Integration 211
6.1 Orientations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
©F
8
6.2.1 Smooth Invariance of Domain in Rn . . . . . . . . . . . . . . . . . . 220
6.2.2 Manifolds with Boundary . . . . . . . . . . . . . . . . . . . . . . . . 222
6.2.3 The Boundary of a Manifold with Boundary . . . . . . . . . . . . . . 224
6.2.4 Tangent Vectors, Differential Forms, and Orientations . . . . . . . . 224
6.2.5 Outward-Pointing Vector Fields . . . . . . . . . . . . . . . . . . . . . 225
ou
6.2.6 Boundary Orientation . . . . . . . . . . . . . . . . . . . . . . . . . . 226
6.3 Integration on Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
6.3.1 The Riemann Integral of a Function on Rn . . . . . . . . . . . . . . . 228
6.3.2 Integrability Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 230
Zh
6.3.3 The Integral of an n-Form on Rn . . . . . . . . . . . . . . . . . . . . 231
6.3.4 Integral of a Differential Form over a Manifold . . . . . . . . . . . . . 232
6.3.5 Integration over a Zero-Dimensional Manifold . . . . . . . . . . . . . 237
6.3.6 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.3.7 Line Integrals & Green’s Theorem . . . . . . . . . . . . . . . . . . . 239
x
eli
©F
9
©F
10
eli
x
Zh
ou
Chapter 1
ou
Euclidean Spaces
Zh
Our goal in this section is to develop calculus in Rn that is independent of coordinates. This
allows us to transition to the setting of manifolds where there is no global coordinate system.
In keeping with the conventions of differentiaal geometry, the indices on coordinates are
superscripts and not subscripts.
©F
Definition 1.1.1 (C k )
Let k ≥ 0. A function f : U → R is said to be C k at p ∈ U if its parptial derivatives
or all orders j ≤ k exist and are continuous at p.
A vector-valued function f : U → Rm is C k at p ∈ U if all its component functions
are C k at p.
We say f : U → R, is C k on U if it is C k at every point p ∈ U .
11
Definition 1.1.2 (Real-Analytic)
We say the function f is real-analytic at a point p if in some neighborhood of p it is
equal to its Taylor series at p:
X 1 X ∂k
f (x) = f (p) + i1 . . . ∂xik
(p)(xi1 − pi1 ) . . . (xik − pik ).
k≥1
k! i ,...,i
∂x
1 k
ou
Recall that a convergent power series can be differentiated term by term in its domain of
convergence. Hence a real-analytic function is necessarily C ∞ . However, the converse need
not hold.
Example 1.1.1
The function f : R → R given by
Zh
(
exp(−1/x) x > 0
x 7→
0, x=0
is C ∞ on R but f (k) (0) = 0 for all k. Hence it cannot be real-analytic about the point
x = 0.
Although a C ∞ function need not be equal to its Taylor series, there is a version of Taylor’s
x
theorem for C ∞ functions that is often good enough.
We say S ⊆ Rn is star-shaped with respect to a point p ∈ S if for every x ∈ S, the line
segment [p, x] lies in S.
eli
Lemma 1.1.2
Let f be C ∞ on an open subset U ⊆ Rn that is star-shaped with respect to some
p ∈ U . There are functions g1 , . . . , gn ∈ C ∞ (U ) such that
n
©F
X
f (x) = f (p) + (xi − pi )gi (x)
i=1
∂f
gi (p) = (p).
∂xi
Proof
Since U is star-shaped with respect to p, for any x ∈ U the line segment yt := p + t(x −
p), t ∈ [0, 1] lies in U . Thus f (yt ) is well-defined for all t ∈ [0, 1].
12
By the chain rule,
d X ∂f
f (yt ) = (xi − pi ) i (yt ).
dt i
∂x
Integrating both sides with respect to t from 0 to 1,
Z t
X
i i ∂f
f (x) − f (p) = (x − p ) (yt )dt.
∂xi
ou
i 0
Taking Z t
∂f
gi (x) := (yt )dt
0 ∂xi
suffices. Indeed, gi ∈ C ∞
since f ∈ C ∞
and gi (p) = ∂f
∂xi
(p).
Let n = 1, p = 0. The lemma above states that
Zh
f (x) = f (0) + xg1 (x)
13
1.2.1 The Directional Derivative
ou
If f ∈ C ∞ in a neighborhood of p and v ∈ Tp (Rn ), recall the directional derivative of f in
the direction v at p is defined as
f (c(t)) − f (p) d
Dv f = lim = f (c(t)).
t→0 t dt t=0
Zh
X dci ∂f X ∂f
Dv f = (0) (p) = v i i (p).
i
dt ∂xi i
∂x
for the map that sends a function f to the number Dv f . We often omit the subscript p
for simplicity of the meaning is clear from context. The association v 7→ Dv offers a way
x
to characterize tangent vectors as certain operators on functions. We study this in greater
detail in the next subsections.
eli
1.2.2 Germs of Functions
As long as two functions agree on some neighborhood of a point p, they will have the same
directional derivatives at p. This suggests that we introduce an equivalence relation on the
C ∞ functions defined in some neighborhood of p.
©F
We write Cp∞ (Rn ) or simply Cp∞ if there is no possibility of confusion for the set of all germs
of C ∞ functions on Rn at p.
14
Example 1.2.1
The functions f (x) = 1/1−x with domain R \ {1} and g(x) = k≥0 xk with domain (−1, 1)
P
have the same germ at any point p in the open interval (−1, 1).
Recall the following definition.
ou
An algebra over a field K is a vector space A over K equiped with a multiplication
map
µ:A×A→A
usually written µ(a, b) = a · b such that for all a, b, c ∈ A and r ∈ K,
(i) (a · b) · c = a · (b · c) [associativity]
(ii) (a + b) · c = a · c + b · c and a · (b + c) = a · b + a · c [distributivity]
Zh
(iii) r(a · b) = (ra) · b = a · (rb) [homogeneity]
For each tangent vector v at a point p ∈ Rn , the directional derivative at p gives a map of
real vector spaces
Dv : Cp∞ → R.
15
By the chain rule, Dv is R-linear and satisfies the Leibneiz rule
Dv (f g) = (Dv f )g(p) + f (p)Dv g
since the partial derivatives have these properties.
ou
or a point-derivation of Cp∞ .
Denote the set of all derivations at p by Dp (Rn ). This is a real vector space since the sum
of two derivations at p and a scalar multiple of a derivation at p are again derivations at p.
We know that directional derivatives at p are all derivations at p, so the map φ : Tp (Rn ) →
Dp (Rn ) from tangent vectors to derivations given by
Zh
* +
∂
v 7→ Dv = v, i .
∂x p
Lemma 1.2.2
If D is a point-derivation of Cp∞ , then D(c) = 0 for any constant function c.
Proof
x
By R-linearity, D(c) = cD(1) and it suffices to prove that D(1) = 0. By the Leibniz rule,
Theorem 1.2.3
The linear map φ(v) := Dv is an isomorphism of vector spaces from the tangent space
Tp (Rn ) to the space of point-derivations of Dp (Rn ).
©F
Proof
Injectivity. Suppose Dv ≡ 0 for some v ∈ Tp (Rn ). We claim that v = 0. To see this, apply
Dv to the coordinate function xj to see that
0 = Dv (xj ) = hv, δij i = v j .
16
shaped at p. By Taylor’s theorem with remainder, there are gi ∈ C ∞ (V ) such that
X
f (x) = f (p) + (xi − pi )gi (x)
i
and gi (p) = ∂x
∂f
i (p). Applying D to both sides and recalling D annihilates constant
functions, we get by the Leibniz rule
ou
X X
D(f ) = [D(xi )gi (p) + pi D(gi )] − pi D(gi )
i i
X
i
= D(x )gi (p)
i
X ∂f
= D(xi ) (p).
i
∂xi
Zh
Thus D = Dv for v = hDx1 , . . . , Dxn i.
This theorem shows that one may identify the tangent vectors at p with the derivations at
p. Under the vector space isomorphism Tp (Rn ) ∼= Dp (Rn ), the standard basis correspondes
to the coordinate partial derivatives From now on, we make this identification and write a
tangent vector v = i v i ei as
P
X ∂
v= vi .
i
∂xi p
The vector space Dp (Rn ) of derivations turns out to be more suitable for generalization to
x
manifolds.
n o
Since Tp (Rn ) has basis ∂/∂xi p
, the vector Xp is a linear combination
X ∂
Xp = ai (p)
i
∂xi p
for some p ∈ U and ai (p) ∈ R. Omitting p, we may write X = i ai ∂/∂xi , where the ai ’s are
P
now functions on U . We say that the vector field X is C ∞ on U if the coefficient functions
17
ai ∈ C ∞ (U ). One can identify vector fields on U with column vectors of functions on U
a1
ai i ↔ ... .
X ∂
X=
∂x
i an
ou
This is the same identification from tangent vectors to derivations, except we now allow the
point p to move in U .
The ring of C ∞ functions on an open set U is commonly denoted by C ∞ (U ) or F(U ). The
multiplication of vector fields by functions on U is defined pointwise:
(f X)p := f (p)Xp .
Zh
If X = ai ∂/∂xi is a C ∞ vector field and f ∈ C ∞ (U ), then
P
i
X
fX = (f ai )∂/∂xi
i
is a C ∞ vector field on U . Thus the set of all C ∞ vector fields on U , denoted X(U ), is not
only a vector space over R, but also a module over the ring C ∞ (U ).
If R is a field, then an R-module is precisely a vector space over R. Thus modules generalize
a vector space by allowing scalars over a ring rather than a field.
©F
18
1.2.5 Vector Fields as Derivations
ou
X ∂f
(Xf )(p) = ai (p) (p)
i
∂xi
X ∂f
Xf = ai .
i
∂xi
Zh
C ∞ (U ) → C ∞ (U )
f 7→ Xf.
Note that a derivation at p is not a derivation of the algebra Cp∞ . A derivation at p is a map
from Cp∞ → R, while a derivation is a map from Cp∞ → Cp∞ .
The set of all derivations is closed under addition and scalar multiplication and forms a
vector space denoted by Der(A). From the discussion above, a C ∞ vector field on an open
19
set U gives rise to a derivation of the algebra C ∞ (U ). We therefore have a map
ϕ : X(U ) → Der(C ∞ (U ))
X 7→ (f 7→ Xf ).
Just as tangent vectors at a point p can be identified with point-derivations of Cp∞ , the vector
fields on an open set U can be identified with the derivations of the algebra C ∞ (U ), ie the
ou
map ϕ is an isomorphism of vector spaces. The injectivity property is not too hard to show
but the surjectivity is non-trivial.
Zh
cross product. a key insight of Grassman, the author of the the multivector, is to work in
the dual space of linear functionals. This provides more flexibility than the viewpoint of
tangent vectors.
If V, W are real vector spaces, we denote by Hom(V, W ) the vector space of all linear maps
f : V → W . Recall the dual space V ∨ of V is the vector space of all real-valued linear
functionals on V
x
V ∨ = Hom(V, R).
The elements of V ∨ are known as covectors or 1-covectors on V .
In the rest of this section, V is some finite-dimensional vector space. If {ei : i ∈ [n]} is
eli
some basis for V , then every v ∈ V is a unique linear combination of the basis vectors.
Let αi : V → R denote the i-th coordinate functional that picks out the i-th coordinate,
αi (v) = v i .
Proposition 1.3.1
The functions {αi : i ∈ [n]} form a basis for V ∨ .
©F
Proof
Any f ∈ V ∨ can be expressed as
X
f (v) = f (ei )αi (v).
i
20
We say the αi ’s form the dual basis of the ei ’s.
Corollary 1.3.1.1
dim V ∨ = dim V for any finite-dimensional vector space V .
1.3.2 Permutations
ou
Fix a positive integer k. A permutation of [k] is a bijection σ : A → A. An r-cycle is a
permutation that is cyclic on some r elements while fixing the others. A transposition is a
2-cycle. Two cycles (a1 . . . ar ) and (b1 . . . bs ) are said to be disjoint if the sets {ai } and {bj }
have empty intersection. The product τ σ of two permutations τ, σ of A is the composition
τ ◦ σ. We write Sk to denote the set of all permutations on [k].
Zh
Recall from elementary group theory that any permutation is the product of disjoint cycles.
Moreover, the sign of a permutation, denoted sgn(σ), takes on value ±1 depending on
whether the permutation is the product of even or odd number of transpositions. This
function is well-defined and satisfies
An inversion in a permutation σ is an ordered pair (σ(i), σ(j)) such that i < j but σ(i) >
σ(j). A second way to compute the sign of a permutation is to count the bnumber of
inversions.
x
Proposition 1.3.2
A permutation is even if and only if it has an even number of inversions.
eli
Proof
The proof is algorithmic and is essentially bubble sort.
Example 1.3.3
The dot product on Rn is bilinear.
21
Example 1.3.4
If we view the determinant as a function of the n column vectors of a matrix, then it is
n-linear in Rn .
ou
f (vσ(1) , . . . , vσ(k) ) = f (v1 , . . . , vk )
Zh
f (vσ(1) , . . . , vσ(k) ) = (sgn σ)f (v1 , . . . , vk )
Example 1.3.5
(i) The dot product on Rn is symmetric
(ii) The determinant on Rn is alternating
(iii) The cross product v × w on R3 is alternating
x
Example 1.3.6
For any two linear functionals f, g ∈ V ∨ , the function
22
Thus f is symmetric if and only if σf = f for all σ ∈ Sk and f is alternating if and only if
σf = (sgn σ)f for all σ ∈ Sk .
Lemma 1.3.7
If σ, τ ∈ Sk and f ∈ Lk (V ), then τ (σf ) = (τ σ)f .
ou
Definition 1.3.3 (Left Action)
If G is a group and X an arbitrary set, a map G × X → X written as
(σ, x) 7→ σ · x
is a left action of G on X if
Zh
(i) e · x = x for every x ∈ X, where e is the identity element in G
(ii) τ · (σ · x) = (τ σ) · x for all τ, σ ∈ G and x ∈ X.
Given any k-linear function f on a vector space V , there is a standard trick to make a
symmetric or alternating k-linear function from f .
©F
X
(Sf )(v1 , . . . , vk ) = f (vσ(1) , . . . , vσ(k) )
σ∈Sk
X
=: σf
σ∈Sk
X
(Af )(v1 , . . . , vk ) = (sgn σ)σf.
σ∈Sk
23
Remark 1.3.8 This trick turns up in different places when taking an average over permuta-
tions can “simply” the problem. For example, we can reduce some instances of semidefinite
programs from algebraic graphs to linear programs by taking an average over permutations
from its automorphism group.
ou
Proposition 1.3.9
If f ∈ Lk (V ),
(i) Sf ∈ Lk (V ) is symmetric
(ii) Af ∈ Lk (V ) is alternating
Zh
Proof
We omit the proof of (i) since it follows the same flow as (ii). For τ ∈ Sk ,
X
τ (Af ) := (sgn σ)τ (σf )
σ∈Sk
X
= (sgn σ)(τ σf )
σ∈Sk
X
= (sgn τ )2 (sgn σ)(τ σf )
σ∈Sk
x
X
= (sgn τ ) (sgn τ σ)(τ σf )
σ∈Sk
= (sgn τ )Af.
eli
Lemma 1.3.10
If f ∈ Lk (V ) is alternating, then
Af = (k!)f.
©F
Proof
By computation,
X
Af = (sgn σ)σf
σ∈Sk
X
= (sgn σ)2 f
σ∈Sk
= (k!)f.
24
1.3.6 The Tensor Product
ou
Let {ei : i ∈ [n]} be a basis for a vector space V {αi } the dual basis for V ∨ , and
h, i : V × V → R a bilinear map on V .
Set gij := hei , ej i ∈ R. If v = i v i ei and w = wi ei , we can express h, i in terms of the
P P
tensor product
X
hv, wi = v i wj hei , ej i
Zh
i,j
X
= αi (v)αj (w)gij
i,j
X
= gij (αi ⊗ αj )(v, w).
i,j
Proposition 1.3.12
The tensor product of multi-linear functions is associative since multiplication is asso-
ciative in R.
x
1.3.7 The Wedge Product
k!`! σ∈S
k+`
25
The coefficient 1/k!`! compensates for repetitions in the sum. For every permutation σ ∈ Sk+` ,
there are k! permutations τ ∈ Sk that permute the first k arguments and leave the arguments
of g alone. For every such τ ∈ Sk , the composition στ ∈ Sk+` contributes the same term to
the sum, since f is alternating. This we divide by k! to get rid of the k! repeating terms in
the sum and similarly for `!.
Another way to avoid redundancies in the definition of f ∧ g is to stipulate that in the sum,
σ(1), . . . , σ(k) is in ascending order and also σ(k + 1), . . . , σ(k + `).
ou
Definition 1.3.5 (Shuffle)
A permutation σ ∈ Sk+` is a (k, `)-shuffle if
and
Zh
σ(k + 1) < · · · < σ(k + `).
Example 1.3.14
The wedge product of two covectors f, g ∈ A1 (V ) is given by
By the definition of the wedge product, f ∧g is bilinear in f and g since it is a sum of bilinear
functions.
Proposition 1.3.15
The wedge product is anticommutative: if f ∈ Ak (V ) and g ∈ A` (V ),
©F
f ∧ g = (−1)k` g ∧ f.
Proof
Define τ ∈ Sk as (
i + k, i≤`
τ (i) :=
i − ` ≡ i + k − (k + `), i > `
26
Thus τ is the k-th product of the k + ` cycle and
σ(1) = στ (` + 1)
...
σ(k) = στ (` + k)
σ(k + 1) = στ (1)
ou
...
σ(k + `) = στ (`).
Zh
X
= (sgn σ)f (vστ (`+1) , . . . , vστ (`+k) )g(vστ (1) , . . . , vστ (`) )
σ∈Sk+`
X
= (sgn τ ) (sgn στ )g(vστ (1) , . . . , vστ (`) )f (vστ (`+1) , . . . , vστ (`+k) )
σ∈Sk+`
The statement can then be proven by dividing by k!`! and verifying that sgn τ = (−1)k` .
Indeed, the (k + `)-cycle has sign (−1)k+`−1 . Taking the composition of k of them yields
(−1)k`+k(k−1) = (−1)k` .
x
Corollary 1.3.15.1
If f is a multicovector of odd degree on V , then f ∧ f = 0.
eli
Proof
Let k be the degree of f . By anticommutativity,
2
f ∧ f = (−1)k f ∧ f = −f ∧ f.
Lemma 1.3.16
Suppose f ∈ Lk (V ) and g ∈ L` (V ).
(i) A(A(f ) ⊗ g) = k!A(f ⊗ g)
(ii) A(f ⊗ A(g)) = `!A(f ⊗ g)
ou
Proof
We prove (i) and omit the proof of (ii) as it follows the exact train of thought. By
definition, !
X X
A(A(f ) ⊗ g) = (sgn σ)σ (sgn τ )(τ f ) ⊗ g .
σ∈Sk+` τ ∈Sk
Zh
We can view τ ∈ Sk as a permutation in Sk+` . Hence
X
A(A(f ) ⊗ g) = (sgn σ)(sgn τ )(στ )(f ⊗ g).
σ∈Sk+` ,τ ∈Sk
For each µ ∈ Sk+` and τ ∈ Sk , σ = µτ −1 is the unique element such that µ = στ . Hence
each µ ∈ Sk+` appears once in the double sum for each τ ∈ Sk , and hence k! times in
total. Thus X
A(A(f ) ⊗ g) = k! (sgn µ)µ(f ⊗ g) = k!A(f ⊗ g).
µ∈Sk+`
x
Proposition 1.3.17 (Associativity of the Wedge Product)
Let V be a real vector space and f ∈ Ak (V ), g ∈ A` (V ), h ∈ Am (V ). Then
eli
(f ∧ g) ∧ h = f ∧ (g ∧ h).
Proof
By the definition of the wedge product,
1
(f ∧ g) ∧ h = A((f ∧ g) ⊗ h)
©F
(k + `)!m!
1 1
= A(A(f ⊗ g) ⊗ h)
(k + `)!m! k!`!
(k + `)!
= A((f ⊗ g) ⊗ h) lemma
(k + `)!m!k!`!
1
= A((f ⊗ g) ⊗ h).
k!`!m!
We can also show that that f ∧ (g ∧ h) is also equal to the last term, concluding the proof
28
by the associativity of the tensor product.
Since associativity holds, it is customary to omit the parenthesis in multiple wedge products.
Corollary 1.3.17.1
If fi ∈ Adi (V ), then
ou
f1 ∧ · · · ∧ fr = A(f1 ⊗ · · · ⊗ fr ).
(d1 )! . . . (dr )!
Let [bij ] denote the matrix whose (i, j)-th entry is bij . We have the following proposition.
Zh
(α1 ∧ · · · ∧ αk )(v1 , . . . , vk ) = det α(vj )i .
Proof
We have
= det αi (vj ) .
x
Recall the notation
∞
M
A= Ak
eli
k=0
a = ai 1 + · · · + ai m
©F
of vector spaces over K such that the multiplication map from A sends Ak × A` to
Ak+` .
29
Definition 1.3.7 (Graded Commutative)
A graded algebra A = ⊕∞ k=0 A is said to be graded commutative or anticommutative
k
ab = (−1)k` ba.
ou
Example 1.3.19
The polynomial algebra A = R[x, y] is the graded by degree; Ak consists of all homoge-
neous polynomials of total degree k in the variables x and y.
For a vector space V of finite dimension n, define
∞ n
Zh
M M
A∗ (V ) := Ak (V ) = Ak (V ).
k=0 k=0
Note the second equality is due to linear dependence. With the wedge product of multicov-
ector as multiplication, A∗ (V ) becomes an anticommutative graded algebra, known as the
exterior algebra or the Grassman algebra of multicovectors on the vector space V .
I = (i1 , . . . , ik )
eli
to write eI := (ei1 , . . . , eik ) and αI for αi1 ∧ · · · ∧ αik .
A k-linear function f ∈ Lk (V ) is completely determined by its values on all k-tuples eI . If
f is alternating, it is completely determined by its values on eI with 1 ≤ i1 < · · · < ik ≤ n.
This it suffices to consider eI with I in strictly ascending order.
©F
Lemma 1.3.20
Let e1 , . . . , en be a basis for a vector space V and α1 , . . . , αn its dual basis in V ∨ . If
I, J are strictly ascending multi-indices of length k, then
(
1, I = J
αI (eJ ) = δJI =
0, I 6= J
30
Proof
We have previously shown that
ou
assume that i` < j` . Then i` is not equal to any of j1 , . . . , j` . But then the `-th row of
the matrix [αi (ej )] is all zero and the determinant evaluates to 0.
Proposition 1.3.21
The alternating k-linear functions αI for all strictly ascending multi-indices I, form a
basis for the space Ak (V ).
Zh
Proof
Linear Independence. Suppose I cI αI = 0. Applying both sides to an arbitrary eJ , we
P
get that X X
0= cI αI (eJ ) = cI δJI = cJ .
I I
Corollary 1.3.21.1
If dim V = n, then Ak (V ) has dimension n
.
k
©F
Proof
The number of strictly ascending multi-indices is the number of size k subsets of [n].
Corollary 1.3.21.2
If k > dim V , then Ak (V ) = 0.
31
Proof
In the multi-index I, at least two will be the same, say i = j. But then αi ∧ αj = 0 since
αi , αj are equal covectors of odd degree 1 and so αI = 0 as well.
ou
Proposition 1.3.22 (Characterizations of Alternating Tensors)
Let f ∈ Lk (V ). The following are equivalent.
(i) f is alternating
(ii) f changes signs when two arguments are interchanged
(iii) f changes signs when two successive arguments are interchanged
Zh
(iv) f (v1 , . . . , vk ) = 0 whenever two of its arguments are equal
k
X
βi = aij γ j
j=1
where A = [aij ].
eli
Proof
Using the fact that α ∧ α = 0 for all α ∈ L1 (V ),
k
! k
!
X X
β1 ∧ · · · ∧ βk = a1j1 γ j1 ∧ ··· ∧ akjk γ jk
©F
j1 =1 jk =1
X k
Y
= aiσ(i) γ σ(1) ∧ · · · ∧ γ σ(k)
σ∈Sk i=1
k
XY
= aiσ(i) (sgn σ)γ 1 ∧ · · · ∧ γ k
σ∈Sk i=1
= (det A)γ 1 ∧ · · · ∧ γ k .
32
Theorem 1.3.24 (Transformation Rule for Multi-Covectors)
Let f ∈ Ak (V ) and suppose vectors u1 , . . . , uk , v1 , . . . , vk ∈ V satisfy
k
X
uj = aij vi
i=1
ou
for each j ∈ [k]. Then
Lemma 1.3.25
Let α1 , . . . , αk ∈ A1 (V ). Then α1 ∧ · · · ∧ αk = 0 if and only if α1 , . . . , αk are linearly
independent within the dual space V ∨ .
Zh
The lemma can be shown using the transformation rule.
Proof (Sketch)
If the latter holds, then clearly α ∧ γ = 0.
x
Suppose α ∧ γ = 0 and write γ as a linear combination of wedge products formed by any
basis of V ∨ that includes α. By linear independence, we see that we can factor out a α
term in each non-zero term in the linear combination.
eli
Thus an element of the cotangent space Tp∗ (Rn ) is a covector on the tangent space Tp (Rn ).
The following definition is the analogue to the vector field.
33
Definition 1.4.2 (Differential 1-Form)
A covector field or a differential 1-form on an open subset U ⊆ Rn is a function ω
that assigns to each point p ∈ U a covector ωp ∈ Tp∗ (Rn ),
G
ω:U → Tp∗ (Rn )
p∈U
ou
p 7→ ωp ∈ Tp∗ (Rn )
(df )p (Xp ) = Xp f.
Zh
The directional derivative of a function in the direction of a tangent vector at a point p sets
up a bilinear function
We can think of a tangent vector as a function on the second argument of this pairing:
hXp , ·i. The differential (df )p at p is then a function on the first argument of the pairing:
h·, f i. The value of the differential df at p is also written df |p .
x
Recall the partial derivatives form a basis to the tangent space Tp (Rn ).
Proposition 1.4.1
eli
Let x1 , . . . , xn denote the standard basis on Rn . At every point p ∈ Rn ,
(dx1 )p , . . . , (dxn )p is the dual basis for the cotangent space Tp∗ (Rn ) with respect to the
basis ∂/∂x1 |p , . . . , ∂/∂xn |p for the tangent space Tp (Rn ).
Proof
By definition, !
∂ ∂
(dxi )p = xi = δji .
∂xj p ∂xj p
34
for some aiP(p) ∈ R. As p varies over U , the coefficients ai become functions on U and we can
write ω = i ai dxi . The covector field ω is said to be C ∞ on U if the coefficient functions
ai are all C ∞ on U .
Proposition 1.4.2
If f : U → R is a C ∞ function on an open subset U ⊆ Rn , then
ou
X ∂f
df = i
dxi .
i
∂x
Proof
We know that
X
(df )p = ai (p)(dxi )p
Zh
i
at every p ∈ U for some ai (p). Conclude by evaluating this at the coordinate vector field
∂/∂xj :
∂f ∂
= df
∂xj ∂xj
X
i ∂
= ai dx
i
∂xj
X
= ai δji
i
x
= aj .
The proposition above shows that if f ∈ C ∞ , then the 1-form df is also C ∞ . Moreover, we
see that dxi is nothing more than the coordinate function with respect to the basis ∂/∂xj .
eli
Since A1 (Tp Rn ) = Tp∗ (Rn ), the definition of a k-form generalizes that of a 1-form we have
just seen.
We have shown that one possible basis for Ak (V ) where dim V < ∞, is the wedge product
35
of k-coordinate functions with strictly ascending indices. Thus a basis for Ak (Tp Rn ) is
ou
I
Zh
Denote by Ωk (U ) the vector space of C ∞ k-forms on U . A 0-form on U assigns to each point
p ∈ U an element of A0 (Tp Rn ) = R. Thus Ω0 (U ) = C ∞ (U ).
The wedge product of a k-form ω and an `-form τ on some open U ⊆ Rn is defined pointwise:
(ω ∧ τ )p = ωp ∧ τp
X
ω∧τ = (aI bJ )dxI ∧ dxJ ,
I, J disjoint
x
which shows that the wedge product of two C ∞ forms is C ∞ . Thus the wedge product is a
bilinear map
∧ : Ωk (U ) × Ω` (U ) → Ωk+` (U ).
eli
Recall that the wedge product is anticommutative and associative.
For the case k = 0, the wedge product is just the pointwise multiplication of a C ∞ function
and a C ∞ `-form
(f ∧ ω)p = f (p) ∧ ωp = f (p)ωp .
Thus if f ∈ C ∞ (U ) and ω ∈ Ω` (U ), then f ∧ ω = f ω.
©F
Example 1.4.3
Let x, y, z be the coordinates in R3 . The C ∞ k-forms on R3 are
36
Example 1.4.4
Let x1 , . . . , x4 be the coordinates in R4 and p ∈ R4 . A possible basis for the vector space
A3 (Tp R4 ) is given by
(dx1 ∧ dx2 ∧ dx3 )p , (dx1 ∧ dx2 ∧ dx4 )p , (dx1 ∧ dx3 ∧ dx4 )p , (dx2 ∧ dx3 ∧ dx4 )p .
With the wedge product as multiplication and the degree of a form as the grading, the
ou
direction sum n
M
∗
Ω (U ) = Ωk (U )
k=0
becomes an anticommutative graded algebra over R. Since we can multiply C ∞ k-forms by
C ∞ functions, the set Ωk (U ) is both a vector space over R and a module over C ∞ (U ). Thus
the direct sum Ω∗ (U ) is also a module over the ring C ∞ (U ).
Zh
1.4.3 Differential Forms as Multilinear Functions on Vector Fields
Recall that we write X(U ) to denote the set of all C ∞ vector fields on U . Thus a C ∞ 1-form
gives rise to a map X(U ) → C ∞ (U ).
This function is linear over the ring C ∞ (U ). To see this, consider some f ∈ C ∞ (U ).
(ω(f X))p := ωp (f (p)Xp )
©F
In the notation F(U ) = C ∞ (U ), a 1-form gives rise to a F(U )-linear map X(U ) → F(U ).
Similarly, a k-form ω on U gives rise to a k-linear map over F(U ),
X(U ) × · · · × X(U ) → F(U )
(X1 , . . . Xk ) 7→ ω(X1 , . . . , Xk ).
37
Example 1.4.5
Let ω be a 2-form and τ a 1-form on R3 . If X, Y, Z are vector fields on R3 ,
ou
In order to define the exterior derivative of a C ∞ k-form on an open subset U ⊆ Rn , we first
define it on 0-forms. Recall Ωk (U ) denotes the C ∞ k-forms on U . The exterior derivative
of a function f ∈ C ∞ (U ) is defined to be its differential df ∈ Ω1 (U ). Recall we showed the
coordinate expansion
X ∂f
df = i
dxi .
i
∂x
Zh
Definition 1.4.4P(Exterior Derivative)
For k ≥ 1, if ω = I aI dxI ∈ Ωk (U ) for some aI ∈ C ∞ (U ), define
X
dω := daI ∧ dxI
I
!
X X ∂aI
= dxj ∧ dxI
I j
∂xj
k+1
∈Ω (U ).
x
Example 1.4.6
Let ω = f dx + gdy ∈ Ω2 (R2 ) be a 2-form where f, g ∈ C ∞ (R2 ). We write fx = ∂/∂x and
eli
fy = ∂f /∂y for simplicity. Then
dω = df ∧ dx + dg ∧ dy
= (fx dx + fy dy) ∧ dx + (gx dx + gy dy) ∧ dy
= (gx − fy )dx ∧ dy.
©F
Furthermore, if there is some m such that the antiderivation D sends Ak → Ak+m for all k,
then we say that it is an antiderivation of degree m.
38
By defining Ak := 0 for k < ∞, we can extend the grading of a graded algebra A to negative
integers. With this extension, the degree m of an antiderivation can be negative.
Proposition 1.4.7
(i) The exterior differentiation d : Ω∗ (U ) → Ω∗ (U ) is an antiderivation of degree 1:
ou
(ii) d2 = 0
(iii) If f ∈ C ∞ (U ) and X ∈ X(U ), then (df )(X) = Xf .
Proof
(i) Since both sides of the claimed equality is linear in ω, τ , it suffices to check the equality
Zh
for ω = f dxI and τ = gdxJ . Then
Moving the 1-form (∂g/∂xi )dxi across the k-form dxI results in the sign (−1)k as desired.
x
(ii) Again, it suffices to check for ω = f dxI . We have
!
X ∂f
2 I i I
d (f dx ) = d dx ∧ dx
eli
i
∂xi
X ∂ 2f
= j ∂xi
dxj ∧ dxi ∧ dxI .
j,i
∂x
39
Proposition 1.4.8 (Characterization of the Exterior Derivative)
If D : Ω∗ (U ) → Ω∗ (U )
(i) is an antiderivation of degree 1
(ii) satisfies D2 = 0
(iii) satisfies (Df )(X) = X for f ∈ C ∞ (U ) and X ∈ X(U )
then D = d.
ou
Proof
By linearity, it suffices to check that D = d on a basic k-form f dxi1 ∧ · · · ∧ dxik .
(iii) states that Df = df on C ∞ functions f . It follows that Ddxi = DDxi = 0 by (ii). But
then by the antiderivation property and induction, D(dxI ) = D(dxi1 ∧(dxi2 ∧· · ·∧dxik )) =
0.
Zh
Finally, for every k-form f dxI ,
Example 1.4.9
Define a 1-form on R2 − {0} by
1
ω= (−ydx + xdy).
x2 + y 2
40
Then ω is closed since
x y
dω = d ∧ dy − d ∧ dx
x2 + y 2 x2 + y 2
= db ∧ dy − da ∧ dx
= ∂x b(dx ∧ dy) − ∂y a(dy ∧ dx)
y 2 − x2 x2 − y 2
ou
= 2 (dx ∧ dy) + (dx ∧ dy)
(x + y 2 )2 (x2 + y 2 )2
= 0.
Zh
For any open subset U ⊆ Rn , the exterior derivative d makes the vector space Ω∗ (U ) of C ∞
forms on U into a cochain complex, called the de Rham complex of U :
d d
0 → Ω0 (U ) −
→ Ω1 (U ) −
→ ....
The closed forms are precisely the elements of the kernel of d and the exact forms are the
elements of the image of d.
P ∂x P Ry − Qz
curl Q = ∂y × Q = −(Rx − Pz )
R ∂z R Qx − Py
P ∂x P
div Q = ∂y · Q = Px + Qy + Rz .
R ∂z R
Now, every 1-form on U is a linear combination with function coefficients of dx, dy, dz. Thus
41
we can identify 1-forms with vector fields on U via
P
P dx + Qdy + Rdz ↔ Q .
R
ou
P
P dy ∧ dz + Qdz ∧ dx + Rdx ∧ dy ↔ Q .
R
f dx ∧ dy ∧ dz ↔ f.
Zh
Following these identifications, the exterior derivative of a 0-form f is
fx
df = fx dx + fy dy + fz dz ↔ fy = grad f.
fz
l
P
Px + Qy + Rz = div Q .
R
In summary, the exterior derivative on 0-forms, 1-forms, and 2-forms are simply the operators
grad, curl, and div, under the appropriate identifications. Under these identifications, a
vector field hP, Q, Ri on R3 is the gradient of a C ∞ function f if and only if the corresponding
1-form P dx + Qdy + Rdz = df .
42
We now state three basic facts concerning grad, curl, and div.
Proposition 1.4.10
The following hold:
(a) curl(grad f ) = h0, 0, 0i
(b) div(curlhP, Q, Ri) = 0
ou
This proposition expresses the property d2 = 0 on open subsets of R3 .
Proposition 1.4.11
On R3 , a vector field F is the gradient of some scalar function f if and only if curl F = 0.
This proposition expresses the fact that a 1-form on R3 is exact if and only if it is closed.
This need not hold on a region other than R3 , as the following examplle shows.
Zh
Example 1.4.12
Define U := R3 − {(0, 0, z) : z ∈ R} and consider the familiar vector field
−y x
F = , ,0
x2 + y 2 x2 + y 2
{closed k-forms on U }
H k (U ) :=
{exact k-forms on U }
©F
called the k-th de Rham cohomology of U . The generalization of the previous proposition
to any differential form on Rn is called the Poincaré lemma: for k ≥ 1, every closed k-form
on Rn is exact. This is equivalent to proving the vanishing of the k-th deRham cohomology
H k (Rn ) for k ≥ 1.
The theory of differential forms allows us to generalize vector calculus from R3 to Rn and
then manifolds of any dimension. The general Stokes theorem subsumes and unifies the
fundamental theorem for line integrals, Green’s theorem in the plane, the classical Stokes
theorem for a surface in R3 , and the divergence theorem.
43
1.4.7 Convention on Subscripts and Superscripts
ou
Coefficient functions may have superscripts or subscripts depending on whether they P are
the coefficients of a vector field or a differential forms. For a vector field X = i a ei ,
i
This
P convention conveniently leads to a “conservation of indices”. For example, if X =
a ∂/∂x i
, then
Zh
i i
ai = (dxi )(X)
so both sides of the equality have a net superscript i. If ω = j bj dxj ,
P
! !
X
j
X
i ∂ X
ω(X) = bj dx a = b i ai .
j i
∂xi i
After cancellation of superscripts and subscripts, both sides of the quality sign have zero net
index. This convention is a useful mnemonic aid in some of the transformation formulas in
differential geometry.
x
1.4.8 Miscellaneous Results
eli
Definition 1.4.9 (Superderivation)
Let A = ⊕∞ k=−∞ A be a graded algebra over a field K with A = 0 for k < 0 and
k k
m ∈ Z.
A superderivation of A of degree m is a K-linear map D : A → A such that for all k,
D(Ak ) ⊆ Ak+m and for all a ∈ Ak , b ∈ A` ,
©F
Proposition 1.4.13
If D1 , D2 are two superderivations of A of respective degrees m1 , m2 , their commutator
[D1 , D2 ] := D1 ◦ D2 − (−1)m1 m2 D2 ◦ D1
is a superderivation of degree m1 + m2 .
44
This proposition can be verified by checking the definitions.
A super derivation is said to be even or odd depending on the parity of its degree. An even
superderivation is a derivation is a derivation and an odd superderivation is an antiderivation.
ou
Zh
x
eli
©F
45
©F
46
eli
x
Zh
ou
Chapter 2
ou
Manifolds
Zh
Intuitively, a manifold is a generalization of curves and surfaces to higher dimension. It is
locally Euclidean in the sense that every point has a neighborhood, called a chart, that is
homeomorphic to an open subset of Rn . The coordinates on a chart allow one to carry out
computations as though in a Euclidean space, so may concepts from Rn , such as differentia-
bility, point-derivations, tangent spaces, and differential forms, carry over to a manifold.
Our goal is to define and explore basic properties of a smooth manifold and smooth maps
between manifolds. Initially, the only way to verify that a space is a manifold is to exhibit a
collection of C ∞ compatible charts covering the space. We eventually see a set of sufficient
conditions under which a quotient topological space becomes a manifold.
x
2.1 Manifolds
eli
2.1.1 Topological Manifolds
We recall a few definitions from point-set topology. A topological space is second countable if
it has a countable basis. It is said to be Hausdorff if any two distinct points are repectively
contained in disjoint open neighborhoods.
©F
We call the pair (U, φ) a chart, U a coordinate neighborhood or a coordinate open set, and φ
a coordinate map or a coordinate system on U . Moreover, we say a chart (U, φ) is centered
47
at p ∈ U if φ(p) = 0.
Remark 2.1.1 We require manifolds to be Hausdorff and second countable to ensure that
ou
manifolds behave as expected from our experience with Euclidean spaces. For instance, finite
subsets are closed and limits of sequences are unique in Hausdorff spaces. The motivation
for second-countability is based on the existence of the so called partitions of unity.
Remark 2.1.2 We can restrict our definition of manifolds by forcing each coordinate map
φ : U → Rn to be homeomorphic to an open ball in Rn or to Rn itself.
Zh
We say a manifold is of dimension n or is an n-manifold if it is locally Euclidean of dimension
n. For the dimension of a manifold to be well-defined, we need to know that for n 6= m,
an open subset of Rn is not homeomorphic to an open subset of Rm . This is known as the
invariance of dimension but is not easy to prove directly. Since we are interested in smooth
manifolds, the analogous result is much easier to prove.
Example 2.1.3
The Euclidean space Rn is covered by a single chart (Rn , IdRn ) and is hence an n-manifold.
x
Similarly, every open subset of Rn is also an n-manifold.
Let (X, τ ) be a topological space and A ⊆ X. Recall that A endowed with the subspace
topology {U ∩ A : U ∈ τ } is called a subspace of X. The Hausdorff condition and second
eli
countability are inherited by subspaces. Thus any subspace of Rn is automatically Hausdorff
and second countable.
Example 2.1.4
The graph of a contiuous function f : R → R is a 1-manifold since it is a subspace of R2
and is locally Euclidean through the homeomorphism (x, f (x)) 7→ x.
©F
Example 2.1.5
The union of the x and y-axis in R2 is not a topological manifold.
Suppose towards a contradition that there is a neighborhood of the intersection p that is
homeomorphic to an open ball B of Rn with p mapping to 0. The homeomorphism U → B
restricts to a homeomorphism U − {p} → B − {0}. Now, B − {0} is either connected for
n ≥ 2 or disconnected if n = 1. Since U − {p} has four connected components, this is a
contradiction.
48
2.1.2 Compatible Charts
Suppose (U, φ), (V, ψ) are two charts of an n-manifold. Since U ∩ V is open in U and
φ : U → Rn is a homeomorphism onto a open subset of Rn , the image φ(U ∩ V ) will also be
an open subset of Rn . Similarly, ψ(U ∩ V ) is an open subset of Rn .
ou
Definition 2.1.3 (Smoothly Compatible)
Two charts (U, φ), (V, ψ) of a topological manifold are C ∞ -compatible if the two maps
φ ◦ ψ −1 : ψ(U ∩ V ) → φ(U ∩ V )
ψ ◦ φ−1 : φ(U ∩ V ) → ψ(U ∩ V )
Zh
These compositions are called transition functions between the charts. We often omit
“smoothly” and simply speak of compatible charts.
If U ∩V is empty, then the two charts are automatically smoothly compatible. For simplicity
of notation, we will sometimes write Uαβ for Uα ∩ Uβ and similarly Uαβγ = Uα ∩ Uβ ∩ Uγ .
φ1 (eit ) = t t ∈ (−π, π)
φ2 (eit ) = t t ∈ (0, 2π)
49
Indeed, Suppose (U1 , φ1 ) is compatible with (U2 , φ2 ) and (U2 , φ2 ) is compatible with (U3 , φ3 ).
Note that the three coordinate functions are simultaneously defined on the intersection U123 .
Thus the composite
φ3 ◦ φ−1 −1 −1
1 = (φ3 ◦ φ2 ) ◦ (φ2 ◦ φ1 )
is smooth but only on φ1 (U123 ) and not necessarily on φ1 (U13 ). Here the equality means on
the restriction to their common domains.
ou
Definition 2.1.5
A chart (V, ψ) is compatible with an atlas U if it is compatible with all the charts of
the atlas.
Lemma 2.1.7
Let U = {(Uα , φα )} be an atlas on a locally Euclidean space. If two charts
Zh
(V, ψ), (W, σ) are both compatible with the atlas, then they are compatible with
each other.
Proof
Let p ∈ V ∩ W . We need to show that σ ◦ ψ −1 is C ∞ at ψ(p). Since U is an atlas for M ,
we can choose Uα 3 p for some α. Then p is in the intersection V ∩ W ∩ Uα .
From the remark above,
σ ◦ ψ −1 = (σ ◦ φ−1 −1
α ) ◦ (φα ◦ ψ )
x
is C ∞ on ψ(V ∩ W ∩ Uα ), and hence at ψ(p). But p ∈ V ∩ W was arbitrary, hence σ ◦ ψ −1
is C ∞ on ψ(V ∩ W ). Similarly, ψ ◦ σ −1 is C ∞ on σ(V ∩ W ).
eli
2.1.3 Smooth Manifolds
the maximal atlas is also called a differentiable structure on M . A manifold is said to have
dimension n if all of its connected components have dimension n. A 1-dimensional manifold
is also called a curve, a 2-dimensional manifold a surface, and an n-dimensional manifold an
n-manifold as before.
In practice, we need not check that a topological manifold M has a maximal atlas. The
50
existence of any atlas on M will do, due to the following proposition.
Proposition 2.1.8
Any atlas U on a locally Euclidean space is contained in a unique maximal atlas.
Proof
Adjoin to the atlas U all charts that are compatible with U. But then add the added
ou
charts are compartible with each other by the previous proposition. Thus the enlarged
collection of charts is still an atlas. Any chart compatible with the new atlas must be
compatible with the original atlas U and so by construction belongs to the new atlas. This
proves that the new atlas is maximal.
Uniqueness follows by adding all charts from another maximal atlas and noting we have
yet another atlas.
Zh
In summary, to show that a topological space M is a smooth manifold, it suffices to check
(i) M is Hausdorff and second countable
(ii) M has a (not necessarily smooth) atlas.
From hereonforth, “manifold” will mean a smooth manifold. In the context of manifolds,
we denote the standard coordinates on Rn by r1 , . . . , rn . If (U, φ : U → Rn ) is a chart of
some manifold, we let xi := ri ◦ φ be the i-th component of φ and write φ = (x1 , . . . , xn ) and
(U, φ) = (U, x1 , . . . , xn ). Thus for p ∈ U , (x1 (p), . . . , xn (p)) is a point in Rn . The functions
are called (local) coordinates on U . By abuse of notation, we sometimes omit the p so the
notation (x1 , . . . , xn ) standard for local coordinates on the open set U and for a point in
x
Rn . By a chart (U, φ) about p in a manifold M , we will mean a chart in the differentiable
structure of M such that p ∈ U .
eli
2.1.4 Examples of Smooth Manifolds
51
countability, this discrete set must be countable.
ou
φ : Γ(f ) → U (x, f (x)) 7→ x
(1, f ) : U → Γ(f ) x 7→ (x, f (x))
Zh
a smooth manifold.
This shows that many of the familiar surfaces of calculus, for example an illiptic paraboloid
or a hyperbolic paraboloid, are manifolds.
52
Proposition 2.1.16
Let {(Uα , φα )} and {(Vi , ψi )} be smooth atlases for the manifolds M, N of dimensions
m, n respectively. Then the collection
{(Uα × Vi , φα × ψi : Uα × Vi → Rm × Rn )}
ou
Example 2.1.17
The infinite cylinder S 1 × R and the torus S 1 × S 1 are manifolds. Moreover, the n-
dimensional torus ×ni=1 S 1 is a manifold.
Remark 2.1.18 Let S n denote the unit sphere
(x1 )2 + . . . (xn+1 )2 = 1
Zh
in Rn+1 . It is not hard to write down a smooth atlas of size 2n on S n by considering
overlapping semi-spheres. The manifold S n with this differentiable structure is called the
standard n-sphere.
One of the most surprising achievements in topology was John Milnor’s discovery of exotic
7-spheres, smooth manifolds homeomorphic but not diffeomorphic to the standard 7-sphere.
In 1963, Michel Kervaire and John Milnor determined that there are exactly 28 nondiffeo-
morphic differentiable structures on S 7 .
x
It is known that in dimensions < 4, every topological manifold has a unique differentiable
structure and in dimensions > 4, every compact topological manifold has a finite number of
differentiable structures. Dimension 4 is a mystery and the statement that S 4 has a unique
differentiable structure is called the smooth Poincaré conjecture.
eli
Michel Kervaire was the first to construct an example of topological manifolds with no
differentiable structure.
Having defined smooth manifolds, we now consider maps between them. Through coordinate
charts, we can transfer the notion of smooth maps from Euclidean spaces to manifolds. It
turns out that smooth compatibility of charts in an atlas means the smoothness of a map is
independent of the choice of charts and is therefore well-defined. We give various criteria for
the smoothness of a map and provide examples of such maps.
Then we transfer the notion of partial derivatives from Euclidean space to a coordinate chart
on a manifold. This enables us to generalize the inverse function theorem to manifolds. Using
53
this result, we formulate a criterion for a set of smooth functions to serve as local coordinates
near a point.
ou
Let M be a smooth n-manifold. A function f : M → R is C ∞ /smooth a a point
p ∈ M if there is a chart (U, φ) about p such that f ◦ φ−1 : φ(U ) → R is C ∞ at φ(p).
Zh
compatibility of the atlas,
f ◦ ψ −1 = (f ◦ φ−1 ) ◦ (φ ◦ ψ −1 )
is smooth at ψ(p). Note we may need to restrict to a smaller neighborhood about p but this
does not change the result.
f = (f ◦ φ−1 ) ◦ φ
x
is a composition of continuous functions at p and is therefore continuous at p.
Since we are only interested in smooth functions on an open set, there is no loss of generality
in assuming at the outset that f is continuous.
eli
(ii) The manifold m has an atlas such that for every chart (U, φ) in the atlas, f ◦ φ−1 :
φ(U ) → R is C ∞
(iii) For every chart (V, ψ) on M , the function f ◦ ψ −1 : ψ(V ) → R is C ∞
The idea of this proposition will be a recurrent motif: to prove the smoothness of an object,
it suffices to show a smoothness criterion holds on the charts of some atlas. Once the object
is shown to be smooth, it then follows that the same smoothness criterion holds on every
chart of the manifold.
54
Definition 2.2.2 (Pullback)
Let F : N → M be a map and h a function on M . The pullback of h by F , denoted
F ∗ h, is the composition h ◦ F .
Thus a function f : M → R is smooth on a chart (U, φ) if and only if its pullback (φ−1 )∗ f
by φ−1 is smooth on φ(U ).
ou
2.2.2 Smooth Maps between Manifolds
Zh
(U, φ) about p ∈ N such that the composition
ψ ◦ F ◦ φ−1 : φ(F −1 (V ) ∩ U ) → Rm
is C ∞ at φ(p).
Proof
©F
Since F is smooth at p ∈ N , there are charts (Uα , φα ) about p ∈ N and (Vβ , ψβ ) about
F (p) ∈ M such that ψβ ◦ F ◦ φ−1
α is smooth at φα (p).
By the smooth compatility of charts in an atlas, both φα ◦ φ−1 and ψ ◦ ψβ−1 are smooth
on open subsets of Euclidean space. Hence the composition
is C ∞ at φ(p).
55
Next, we present a way to check smoothness of a map without specifying points in the
domain.
Proposition 2.2.6
Let N, M be smooth manifolds and F : N → M be continuous. The following are
equivalent:
(i) F is C ∞
ou
(ii) There are atlases U for N and B for M such that for every chart (U, φ) in U and
(V, ψ) in B, the following map is C ∞
ψ ◦ F ◦ φ−1 : φ(U ∩ F −1 (V )) → Rm
(iii) For every chart (U, φ) on N and (V, ψ) on M , the following map is C ∞
Zh
ψ ◦ F ◦ φ−1 : φ(U ∩ F −1 (V )) → Rm
Proof
Let (U, φ), (V, ψ) and (W, σ) be charts on N, M, P respectively. Then
σ ◦ (G ◦ F ) ◦ φ−1 = (σ ◦ G ◦ ψ −1 ) ◦ (ψ ◦ F ◦ φ−1 )
x
under a suitable restriction. By a previous proposition, this is then a composition of
smooth maps on Euclidean space and is therefore smooth. The same proposition now
ensures G ◦ F is smooth. Note we need to let ψ vary over all charts for M to check the
eli
composition is smooth over its entire domain.
2.2.3 Diffeomorphisms
©F
The next two propositions hsow that coordinate maps are diffeomorphisms and conversely,
every diffeomorphism of an open subset of a manifold with an open subset of a Euclidean
space can serve as a coordinate map.
56
Proposition 2.2.8
If (U, φ) is a chart on an n-manifold M , the coordinate map φ : U → φ(U ) ⊆ Rn is a
diffeomorphism.
Proof
φ is a homeomorphism by definition so it suffices to check that φ, φ−1 are both smooth.
ou
We use a previous proposition to check that φ : U → φ(U ) is a smooth map between the
open submanifolds U, φ(U ) of M, Rn respectively. It suffices to show that φ is smooth
with respect to particular atlases for U, φ(U ). Indeed, consider the single chart atlases
{(U, φ)} and {(φ(U ), Id)}. We see that
Id ◦φ ◦ φ−1 = Id
is trivially C ∞ . The same atlases can be used to show that φ−1 is also C ∞ .
Zh
Proposition 2.2.9
Let U ⊆ M be an open subset of the n-manifold M . If F : U → F (U ) ⊆ Rn is a
diffeomorphism onto an open subset of Rn , then (U, F ) is a chart in the differentiable
structure of M .
Proof
For any chart (Uα , φα ) in the maximal atlas of M , both φα , φ−1
α are C
∞
by the previous
proposition. As compositions of smooth maps, both F ◦ φ−1 α and φα ◦ F −1
are C ∞ . Hence
(U, F ) is compatible with the maximal atlas so that (U, F ) must belong in the atlas by
x
maximality.
eli
We now derive a criterion that reduces the smoothness of a map to the smoothness of real-
valued functions on open sets.
©F
57
Proposition 2.2.11 (Smoothness in terms of Components)
Let N be a manifold. A vector-valued function F : N → Rm is C ∞ if and only if its
component functions F 1 , . . . , F m : N → R are all C ∞ .
Example 2.2.12
The map F : R → S 1 given by
ou
F (t) = (cos t, sin t)
is C ∞ .
Zh
(i) F is C ∞
(ii) M has an atlas such that for every chart (V, ψ) = (V, y 1 , . . . , y m ) in the atlas, the
vector-valued function ψ ◦ F : F −1 (V ) → Rm is smooth
(iii) For every chart (V, ψ) = (V, y 1 , . . . , y m ) on M , the vector-valued function ψ ◦ F :
F −1 (V ) → Rm is C ∞
This smoothness criterion then translates into a smoothness criterion in terms of the com-
ponents of the map.
π(p, q) = p
58
Proof
Fix (p, q) ∈ M × N and let (U, φ) = (U, x1 , . . . , xm ) and (V, ψ) = (V, y 1 , . . . , y m ) be
coordinate neighborhoods of p, q respectively. We know that (U × V, φ × ψ) = (U ×
V, x1 , . . . , xm , y 1 , . . . , y n ) is a coordinate neighborhood of (p, q). But then
ou
is a smooth map from (φ × ψ)(U × V ) ⊆ Rm+n to φ(U ) ⊆ Rm . Thus π is smooth at (p, q)
and by the arbitrary choice of (p, q), π is C ∞ on M × N .
Proposition 2.2.16
Let M1 , M2 , N be manifolds of dimension m1 , m2 , n respectively. A map (f1 , f2 ) : N →
M1 × M2 is smooth if and only if f1 , f2 are both smooth.
Zh
Proof
If (f1 , f2 ) is smooth, then the projection is smooth by the previous proposition.
Conversely, suppose f1 , f2 are smooth. Then the components of f1 , f2 on any local coordi-
nate neighborhoods of M1 , M2 respectively are smooth. But there is an atlas U of M1 ×M2
which is just the cross product of charts from M1 , M2 and the components of (f1 , f2 ) on
the local coordinate neighborhoods of U are smooth by assumption. This suffices to show
that (f1 , f2 ) is indeed smooth.
Proposition 2.2.17
A smooth function f (x, y) on R2 restricts to a smooth function on S 1 .
x
Proof (Sketch)
We denote a point on S 1 as p = (a, b) and use x, y to mean the standard coordinate
eli
functions on R2 . If we show that x, y restricts to C ∞ functions on S 1 , then the inclusion
map i(p) = (x(p), y(p)) is then smooth on S 1 and the composition f |S 1 = f ◦ i will
therefore be smooth.
We can check that x, y are smooth on a particular atlas.
The definition of a smooth map between manifolds allows us to define a Lie group.
©F
59
Definition 2.2.5 (Lie Group)
A Lie group is a smooth manifold G with a group structure such that the multiplica-
tion map
µ:G×G→G
and inverse map
ι:G→G
ou
are both smooth.
We can similarly define a topological group which is a topological space with a group structure
wher the multiplication and inverse maps are both continuous. Note a topological group is
not required to the a topological manifold.
Example 2.2.18
Zh
(i) Rn is a Lie group under addition
(ii) C× = C \ {0} is a Lie group under multiplication
(iii) The unit circle S 1 ⊆ C× is a Lie group under multiplication
(iv) The Cartesian product G1 × G2 of two Lie groups (G1 , µ1 ) and (G2 , µ2 ) is a Lie
group under coordinatewise multiplicative µ1 × µ2
60
2.2.6 Partial Derivatives
Definition 2.2.6
Let (U, φ) = (U, x1 , . . . , xn ) be a chart on an n-manifold M , r1 , . . . , rn the standard
coordinates on Rn , and f : U → R a smooth function.
The partial derivative ∂f /∂xi of f with respect to xi at p is given by
ou
∂ ∂f
f := (p)
∂xi p ∂xi
∂(f ◦ φ−1 )
:= (φ(p))
∂ri
∂
:= (f ◦ φ−1 ).
∂ri φ(p)
Zh
The partial derivative ∂f /∂xi is C ∞ on U because its pullback (∂f /∂xi ) ◦ φ−1 is C ∞ on
φ(U ).
The next proposition states that partial derivatives on a manifold satisfy the same duality
property ∂ri /∂rj = δji as the coordinate functions on Rn .
Proposition 2.2.20
Suppose (U, x1 , . . . , xn ) is a chart on a manifold. Then ∂xi /∂xj = δji .
x
Proof
At a point p ∈ U , by the definition of ∂/∂xj |p ,
61
If M, N have the same dimension, then det[∂F i /∂xj ] is known as the Jacobian deter-
minant of F relative to the two charts. The Jacobian determinant is also written as
∂(F 1 , . . . , F n )/∂(x1 , . . . , xn ).
ou
subsets of Rn . Its Jacobian matrix J(ψ ◦ φ−1 ) at φ(p) is the matrix [∂y i /∂xj ] of partial
derivatives at p.
Indeed,
Zh
= (φ(p))
∂rj
∂y i
= (p).
∂xj
62
Theorem 2.2.23 (Inverse Function Theorem for Manifolds)
Let F : N → M be a smooth map between two manifolds of the same dimension,
and p ∈ N . Suppose for some charts (U, φ) = (U, x1 , . . . , xn ) about p ∈ N and
(V, ψ) = (V, y 1 , . . . , y n ) about F (p) ∈ M we have F (U ) ⊆ V . Set F i := y i ◦ F . Then
F is locally invertible at p if and only if its Jacobian determinant det[∂F i /∂xj (p)] is
non-zero.
ou
Proof
Since F i = y i ◦ F = ri ◦ ψ ◦ F , the Jacobian matrix of F relative to the charts (U, φ) and
(V, ψ) is
∂(ri ◦ ψ ◦ F ◦ φ−1 )
i
∂(ri ◦ ψ ◦ F )
∂F
(p) = (p) = (φ(p)) ,
∂xj ∂xj ∂rj
which is precisely the Jacobian matrix at φ(p) of the map
Zh
ψ ◦ F ◦ φ−1 : φ(U ) → ψ(V )
∂r ◦ (ψ ◦ F ◦ φ−1 )
i i
∂F
det (p) = det (φ(p)) 6= 0
∂xj ∂rj
if and only if ψ ◦ F ◦ φ−1 is locally invertible at φ(p). Since ψ, φ are diffeomorphisms, this
x
last statement is equivalent to the local invertibility of F at p.
We typically apply the inverse function theorem in the following form.
Corollary 2.2.23.1
eli
Let N be an n-manifold. A set of n smooth functions F 1 , . . . , F n defined on a coordinate
neighborhood (U, x1 , . . . , xn ) of a point p ∈ N forms a coordinate system aout p if and
only if the Jacobian determinant det[∂F i /∂xj (p)] is non-zero.
Example 2.2.24
©F
63
2.3 Quotients
Recall that given an equivalence relation on a topological space, we can always imbue the
quotient space with a topology such that the natural projection map is continuous. However,
even if the original space is a mnaifold, a quotient space is often not a manifold. We study
conditions under which a quotient space remains second countable and Hausdorff.
ou
2.3.1 The Quotient Topology
Zh
If S is a topological space, we can define a topology on S/∼ by declaring a set U ⊆ S/∼
to be open if and only if π −1 (U ) is open in S. Note the projection map is automatically
continuous by definition under this topology. We call this the quotient topology on S/∼ and
under this topology, S/ ∼ is called the quotient space.
Let ∼ be an equivalence relation on the topological space S and give S/∼ the quotient
topology. Suppose f : S → Y takes values in another topology space Y is constant on each
x
equivalence class. Then it induces a map f¯ : S/∼ → Y given by
[p] 7→ f (p).
eli
Proposition 2.3.1
The induced map f¯ : S/∼ → Y is continuous if and only if the map f : S → Y is
continuous.
This proposition gives a useful criterion for verifying continuous of f¯: lift the function to
f¯ = f ◦ π on S and check the continuous of the lifted map f . The proof of the proposition
©F
64
Example 2.3.2
Let I = [0, 1] be the closed unit interval and I/∼ the quotient space obtained from I
by identifying {0, 1}. Denote by S 1 the unit circle in the complex plane. The function
f : I → S 1 given by
x 7→ exp(2πix)
assumes the same value at 0 and 1 and so induces a function f¯ : I/∼ → S 1 .
ou
Proposition 2.3.3
The function f¯ : I/∼ → S 1 is a homeomorphism.
Proof
Since f is continuous, f¯ is also continuous. Clearly, f¯ is a bijection. As the continuous
image of the compact set I, the quotient I/∼ is compact. Thus f¯ is a continuous bijection
Zh
from a compact space to a Hausdorff space.
From elementary point-set topology, the continuous image of compact (and thus closed)
spaces are again compact (and thus closed) so f¯ is a closed mapping. But then the inverse
is continuous since the pre-image of closed sets are closed. Hence f¯ is a homeomorphism.
The quotient construction does not in general preserve the Hausdorff or second countability
properties. We can derive a necessary condition through the following observation: Every
x
singleton in a Hausdorff space is closed. Thus if π : S → S/∼ is the projection and the
quotient is Hausdorff, then for any p ∈ S, its image {π(p)} is closed in S/∼. By the
continuity of π, the inverse image π −1 (π(p)) = [p] is closed in S.
eli
Proposition 2.3.4
If the quotient space S/∼ is Hausdorff, then the equivalence class [p] of any point p ∈ S
is necessarily closed in S.
Example 2.3.5
©F
We derive conditions under which a quotient space is Hausdorff or second countable. Recall
that a map f : X → Y of topological spaces is an open mapping if the image of open sets
under f is open.
65
Definition 2.3.1 (Open Equivalence Relation)
An equivalence relation ∼ on a topological space S is said to be open if the projection
map π : S → S/∼ is open.
In other words, ∼ on S is open if and only if for every open set U ⊆ S, the set
[
π −1 (π(U )) =
ou
[x]
x∈U
Example 2.3.6
Consider R with ∼ the equivalence relation that identifies 1, −1. Then
Zh
is not open in R so ∼ cannot be open.
Given an equivalence relation ∼ on S, let R ⊆ S × S be the subset that defines the relation
R := {(x, y) ∈ S × S : x ∼ y}.
We say that R is the graph of ∼.
Theorem 2.3.7
Let ∼ be an open equivalence relation on a topological space S. Then the quotient
x
space S/∼ is Hausdorff if and only if the graph R of ∼ is closed in S × S.
Proof
eli
By definition, R is closed in S × S if and only if (S × S) − R is open in S × S. In other
words, for every (x, y) ∈ S × S − R, there is a basic open set U × V containing (x, y)
such that U × V ⊆ S × S − R. This can be restated as for every x y ∈ S, there are
neighborhoods U 3 x, V 3 y in S such that no element of U is equivalent to an element
of V . This is equivalent to the statement that for any two points [x] 6= [y] ∈ S/∼, there
are neighborhoods U 3 x, V 3 y in S such that π(U ) ∩ π(V ) = ∅ in S/∼.
©F
66
Corollary 2.3.7.1
A topological space S is Hausdorff if and only if the diagonal ∆ in S × S is closed.
Theorem 2.3.8
Let ∼ be an open equivalence relation on a topological space S with projection π :
S → S/∼. If B := {Bα } is a basis of S, then the image {π(Bα )} under π is a basis
ou
for S/∼.
Proof
Let W ⊆ S/∼ be open and pick [x] ∈ W for some x ∈ S. Then π −1 (W ) 3 x is open in
S and there is some Bα 3 x contained in π −1 (W ). Then [x] = π(x) ∈ π(Bα ) ⊆ W as
desired.
Zh
Corollary 2.3.8.1
If ∼ is an open equivalence relation on a second-countable space S, then the quotient
space S/∼ is second countable.
We denote the equivalence class of a point (a0 , . . . , an ) ∈ Rn+1 − {0} by [a0 , . . . , an ] and let
π : Rn+1 − {0} → RP n be the projection. We call [a0 , . . . , an ] homogeneous coordinates on
RP n .
©F
Geometrically speaking, two nonzero points in Rn+1 are equivalent if and only if they lie on
the same line through the origin, so RP n can be interpreted as the set of all lines through
the origin in Rn+1 . Such a line uniquely determines a pair of antipodal points on S n . This
suggests that we define an equivalence relation on S n by identifying antipodal points:
x ∼ y ⇐⇒ x = ±y.
67
Proposition 2.3.9 (Real Projective Space as a Quotient of a Sphere)
For x = (x1 , . . . , xn ) ∈ Rn , let kxk denote the Euclidean norm of x. The map f :
Rn+1 − {0} → S n given by
x
x 7→
kxk
induces a homeomorphism f¯ : RP n → S n /∼.
ou
Proof
Since f is continuous on Rn+1 − {0}, π∼ ◦ f : Rn+1 − {0} → S n /∼ is a composition
of continuous functions and is therefore continuous. By a previous proposition, π∼ ◦ f is
constant on the equivalence classes of RP n and the natural quotient function f¯ is therefore
also continuous.
Consider the identity map Id : Rn+1 → Rn+1 . This is certainly continuous. But then
Zh
g : S n → Rn+1 /∼ given by g = π∼ ◦ Id ◦ι where ι is the the inclusion map is continuous.
Since g is constant over the equivalence classes of ∼, the natural quotient function ḡ is
also continuous.
By observation,
x
f¯[x , . . . , x ] =
0 n
kxk
ḡ[x] = [x0 , . . . , xn ]
by
(x, y, z) 7→ (x, y)
is a homeomorphism. Let D2 /∼ denote the closed unit disk with antipodal points iden-
tified. ϕ induces a homeomorphism from H 2 /∼ → D2 ∼. Thus RP 2 ∼= D2 /∼.
RP 2 cannot be embedded as a submanifold into R3 . However, if we allow self-intersection,
we can map RP 2 → R3 as a cross-cap. This map is not injective.
68
Proposition 2.3.12
The equivalence relation on Rn+1 − {0} in the definition of RP n is an open equivalence
relation.
Proof
For an open subset U ⊆ Rn+1 − {0}, the image π(U ) is open in RP n if and only if
π −1 (π(U )) is open in Rn+1 − {0}. But π −1 (π(U )) consists of all nonzero scalar multiples
ou
of points of U . That is,
[ [
π −1 (π(U )) = tU = {tp : p ∈ U }.
t∈R×
Zh
Corollary 2.3.12.1
The real projective space RP n is second countable.
Proposition 2.3.13
The real projective space RP n is Hausdorff.
Proof
Let S = Rn+1 − {0} and consider the set
x
R = {(x, y) ∈ S × S : ∃t ∈ R× , y = tx}.
Ui := {[a0 , . . . , an ] : ai 6= 0}
69
for each i ∈ 0, 1, . . . , n. Define φi : Ui → Rn given by
!
a0 abi an
[a0 , . . . , an ] 7→ , . . . , , . . . ,
ai ai ai
where the carat sign indicates the entry is to be omitted. This map has a continuous inverse
ou
(b1 , . . . , bn ) 7→ [b1 , . . . , |{z}
1 , . . . , bn ].
i-th entry
Zh
xi = , i ∈ [n].
a0
and the coordinate functions on U1 as y 1 , . . . , y n
a0 i ai
y1 = , y = , i ∈ {2, . . . , n}.
a1 a1
So on U0 ∩ U1 ,
1 x2 xn
(φ1 ◦ φ−1
0 )(x) = , , . . . , .
x1 x1 x1
This is a smooth function since x1 6= 0 on φ0 (U0 ∩ U1 ). A similar formula holds on any other
x
Ui ∩ Uj . Thus the collection {(Ui , φi )}i=0,...,n is a smooth atlas for RP n , called the standard
atlas. We conclude that RP n is a smooth manifold as desired.
eli
2.3.8 The Grassmannian Manifold
Define S/G to be the quotient space, also known as the orbit space of the action. Then
the projection map π : S → S/G is an open map.
Proof
It suffices to show that π −1 (π(U )) is open for every U ⊆ S open.
Note that right multiplication by some g ∈ G is a homeomorphism S → S. We can see
this by considering {g} as a topological space under the subspace topology and using the
70
continuity of the inclusion map.
We can express [
π −1 (π(U )) = Ug
g∈G
ou
k-dimensional subspaces of Rn . Such a subspace is completely determined by a full-rank
matrix A ∈ Rn×k , known as a matrix representative of the k-plane. Two matrices A, B
determine the same k-plane if there is a change of basis matrix g ∈ GL(k, R) such that
B = Ag.
Let F (k, n) be the set of all n × k matrices of rank k, topologized as a subspace of Rn×k . We
define the equivalence relation
Zh
A × B ⇐⇒ ∃g ∈ GL(k, R), B = Ag.
Then there is a bijection between G(k, n) and F (k, n)/∼ and we give G(k, n) the subspace
topology F (k, n)/∼.
Proposition 2.3.15
F (k, n)/∼ is Hausdorff and second-countable.
Proof
We can view F/∼ as the orbit space of the action
x
F (n, k) × GL(k, R) → F (n, k)
(A, g) 7→ Ag
eli
which is certainly a continuous group action. But then ∼ is an open relation by the
previous proposition. It follows immediately that F (n, k)/ ∼ is second countable.
In order to show that F (n, k)/ ∼ is Hausdorff, it suffices to show the graph R ⊆ F (n, k) ×
F (n, k) of the relation ∼ is closed. We have
= {(A, B) : rank A B ≤ k}
= {(A, B) : all (k + 1) × (k + 1) minors have determinant 0}.
The last set is a finite intersection of zero-sets of polynomials, which is certainly closed.
We now construct a smooth atlas for G(n, k).
Let
I := {(i1 , . . . , ik ) : 1 ≤ i1 < · · · < ik ≤ n}
71
be the set of increasing row indices. Write AI to denote the submatrix of A made of the
rows of A indexed by I. For I ∈ I, define
VI := {A ∈ F (k, n) : det AI 6= 0}
A 7→ (AA−1 −1
ou
I )I c = AI c AI
where (·)I c indicates taking the rows not indexed by I. This is certainly a continuous map.
Note that if A ∈ VI , then Ag ∈ VI for any g ∈ GL(k, R) since
Zh
Proposition 2.3.16
Let φI : (UI := F (k, n)/∼) → R(n−k)×k denote the map induced by φ̃I . Then φI is
well-defined and
{(UI , φI )}I∈I
is a smooth structure on G(n, k).
Proof
We first check that φ̃I is constant on the orbits of GL(k, n). Indeed,
(Ag)I c (Ag)−1 −1 −1
I = AI c gg AI
x
= AI c A−1
I .
Then we note that φ̃I : VI → R(n−k)×k is continuous since matrix inversion and multipli-
eli
cation are smooth mappings. But then φI : UI → R(n−k)×k is continuous as desired.
To see that it has a continuous inverse, define B (I) as the matrix obtained from B by
inserting the identity matrix in the I-th rows and consider the mapping
R(n−k)×k → UI
B 7→ [B (I) ].
©F
This is certainly continuous since it is the composition of a continuous map with the
projection map. To see that this is the inverse of φI , observe that every element [A] ∈ UI
has a canonical representative
AA−1I
whose I-th rows form the identity matrix. φ̃ selects the I c rows from the canonical
representation and the map above sends the I c rows back to the equivalence class of the
canonical representation.
72
Finally, we need to show the smooth compatibility of charts. Pick I, J ∈ I and consider
φI ◦ φ−1
J : φJ (UI ∩ UJ ) → φI (UI ∩ UJ )
This is smooth since matrix inversion and multiplication are both smooth mappings.
Thus we see that G(n, k) is a smooth (n − k) × k-manifold.
ou
Zh
x
eli
©F
73
©F
74
eli
x
Zh
ou
Chapter 3
ou
The Tangent Space
Zh
By definition, the tangent space to a manifold at a point is the vector space of derivations at
the point. A smooth map of manifolds induces a linear map between tangent spaces at cor-
responding points, called its differential. In local coordinates, the differential is represented
by the Jacobian of partial derivatives of the map.
A basic principle in manifold theory is the linearization principle, meaning a manifold can
be approximated near a point by its tangent space at the point, and a smooth map can be
approximated by the differential of the map. One example of this is the inverse function
theorem.
x
3.1 The Tangent Space
One can define a tangent vector as an arrow in the image of the chart. However, this approach
eli
is complicated since a different chart would give rise to a different set of tangent vectors.
The cleaner approach is to consider point-derivations.
75
multiplication of functions make Cp∞ (M ) into a ring. Scalar multiplication by reals make
Cp∞ (M ) and algebra over R.
We similarly generalize the definition of a point-derivation from Rn .
ou
Cp∞ (M ) → R such that
Zh
Just as for Rn , the tangent vectors at p form a vector space Tp (M ) = Tp M .
Remark 3.1.1 If U is an open set containing p in M , then the algebra Cp∞ (U ) is the same
as Cp∞ (M ). Hence Tp U = Tp M .
76
Proposition 3.1.2
F∗ (Xp ) is a derivation at F (p) and F∗ : Tp N → TF (p) M is a linear map.
Proof
Let f, g ∈ Cp∞ .
ou
F∗ (Xp )(f g) := Xp (f g ◦ F )
= Xp ((f ◦ F )(g ◦ F ))
= Xp (f ◦ F )(g ◦ F (p)) + (f ◦ F (p))Xp (g ◦ F (p))
= F∗ (Xp )f g(F (p)) + f (F (p))F∗ (Xp )g
as desired.
To see linearity, let α ∈ R and Xp , Yp ∈ Tp N . For any f ∈ CF∞(p) ,
Zh
F∗ (αXp + Yp ) := (αXp + Yp )(f ◦ F )
= αXp (f ◦ F ) + Yp (f ◦ F )
= αF∗ (Xp )f + F∗ (Yp )f.
We sometimes write F∗,p = F∗ to make the dependence on p explicit.
!
∂ ∂ ∂F i
F∗ = (y i ◦ F ) = (p).
∂xj p ∂xj p ∂xj
77
3.1.3 The Chain Rule
ou
Proof
Let Xp ∈ Tp N and f a smooth function at G(F (p)) ∈ P .
((G ◦ F )∗ Xp )f = Xp (f ◦ G ◦ F )
=: (F∗ Xp )(f ◦ G)
Zh
=: (G∗ (F∗ Xp ))f
= ((G∗ ◦ F∗ )Xp )f.
Remark 3.1.5 The differential of the identity map M → M is the identity map Tp M →
Tp M .
Corollary 3.1.5.1
If F : M → M is a diffeomorphism of manifolds, then F∗ : Tp N → TF (p) M is an
isomorphism of vector spaces for any p ∈ N .
x
Proof
By assumption, F has a smooth inverse G : M → N . By the chain rule,
eli
F∗ ◦ G∗ = (F ◦ G)∗
= (IdM )∗
= IdTp M
Proof
Let F : U → V be any diffeomorphism and p ∈ U . The previous corollary informs us
that F∗,p : Tp U → TF (p) is an isomorphism of vector spaces. Since there are vector space
isomorphisms Tp U ∼= Rn and TF (p) ∼ = Rm , we must have n = m.
78
3.1.4 Bases of Tangent Space at a Point
As usual, we denote by {ri } the standard coordinates on Rn and if (U, φ) is a chart about
a point p in a an n-manifold M , we set xi = ri ◦ φ. Since φ : U → Rn is a diffeomorphism
onto its image, a previous corollary tells us that the differential
ou
φ∗ : Tp M → Tφ(p) Rn
Zh
Proposition 3.1.6
Let (U, φ) = (U, x1 , . . . , xn ) be a chart about p ∈ M . Then
!
∂ ∂
φ∗ i
= .
∂x p ∂ri φ(p)
Proof
Let f ∈ Cφ(p)
∞
Rn .
x
!
∂ ∂
φ∗ f= (f ◦ φ)
∂xi p ∂xi p
eli
∂
= (f ◦ φ ◦ φ−1 )
∂ri φ(p)
∂
= f.
∂ri φ(p)
Proposition 3.1.7
©F
Let (U, φ) = (U, x1 , . . . , xn ) be a chart containig p, then the tangent space Tp M has
basis
∂ ∂
1
,..., n .
∂x p ∂x p
Proof
An isomorphism of vector spaces carries a basis to a basis. The image of the tangent
vectors above map to the partial derivatives at φ(p), which forms a basis of Tφ(p) Rn .
79
Proposition 3.1.8 (Transition Matrix for Coordiante Vectors)
Suppose (U, x1 , . . . , xn ) and (V, y 1 , . . . , y n ) are two coordinate charts on a manifold M .
Then on U ∩ V ,
∂ X ∂y i ∂
= .
∂xj i
∂xj ∂y i
ou
Proof
At each p ∈ U ∩ V , there are two bases for the tangent space Tp M and so there is a matrix
[aij (p)] such that
∂ X
k ∂
= a j .
∂xj k
∂y k
Zh
∂y i X ∂y i
j
= akj k
∂x k
∂y
X
= akj δki
k
i
= aj .
x
3.1.5 A Local Expression for the Differential
eli
The partial derivatives at p, F (p) form a basis for Tp N, TF (p) M respectively. Thus the dif-
ferential F∗ = F∗,p is completely determined by the numbers aij
!
∂ X ∂
F∗ = akj .
∂xj p k
∂y k F (p)
80
Applying both sides to y i , we find that
!
X ∂
aij = akj yi
k
∂y k F (p)
!
∂
= F∗ yi
∂xj p
ou
∂
= (y i ◦ F )
∂xj p
i
∂F
= (p).
∂xj
Proposition 3.1.9
Given a smooth map F : N → M of manifolds and a point p ∈ N , let (U, (xj )) and
Zh
(V, (y i )) be coordinate charts about p ∈ N and F (p) ∈ M respectively. Relative to
the bases {∂/∂xj |p } for Tp N and {∂/∂y i |F (p) } for TF (p) M , the differential F∗,p : Tp N →
TF (p) M is represented by the matrix
i
∂F
(p) .
∂xj
Remark 3.1.10 The inverse function theorem for manifolds has a coordinate-free descrip-
tion: A smooth map F : N → M between two manifolds of the same dimension is locally
invertible at p ∈ N if and only if its differential at p F∗,p : Tp N → TF (p) M is an isomorphism.
x
3.1.6 Curves in a Manifold
eli
Definition 3.1.5 (Smooth Curve)
A smooth curve in a manifold M is a smooth map c : (a, b) → M .
Thus c0 (t0 ) is simply the differential of the curve at the point t = t0 . We say that c0 (t0 ) is
the velocity of c at the point c(t0 ).
81
If we need to distinguish between the standard calculus notation c0 (t)and the tangent vector
c0 (t0 ), we will write ċ(t) to denote the calculus derivative.
d
c0 (t0 ) = ċ(t0 )
ou
.
dx c(t0 )
Proof
Pick f ∈ Cc(t
∞
0)
. By the chain rule from calculus,
!
d
c0 (t0 )f := c∗ f
Zh
dt t0
d
:= (f ◦ c)
dt t0
d
= ċ(t0 ) f.
dx c(t0 )
Example 3.1.12
Define c : R → R2 by
x
c(t) := (t2 , t3 ).
Then c0 (t) is a linear combination of ∂/∂x and ∂/∂y at c(t):
eli
d ∂ ∂
c∗ = c0 (t) = a +b .
dt ∂x ∂y
We can evaluate both sides at x, y respectively to extract the coefficients a = 2t, b = 3t2 .
More generally, to compute the velocity vector of a smooth curve c ∈ Rn , one can simply
differentiate the components of c.
©F
82
Proof
We already know that c0 (t) is a linear combination of the partial derivatives at c(t), say
with coefficients ai . Evaluating the tangent vector at xi yields
ai = c0 (t)xi
∂
:= c∗ xi
ou
∂t
∂
= (xi ◦ c)
∂t
∂
= ci
∂t
= ċi (t).
Every smooth curve c at a point p ∈ M gives rise to a tangent vector c0 (0) ∈ Tp M . Conversely,
Zh
we can show that every tangent vector Xp ∈ Tp M is precisely the velocity vector of some
curve at p.
Proof
Let
P (U, φ) = (U, x1 , . . . , xn ) be a chart centered at p, ie φ(p) = 0 ∈ Rn . Suppose Xp =
i a ∂/∂x |p at p. Let {r } be the standard coordinates on R and write x = r ◦ φ.
i i i n i i
x
We wish to find a curve α : (−ε, ε) → Rn such that α(0) = 0 and α0 (0) = i ai ∂/∂ri |0 .
P
The previous proposition states that one possible choice of α is given by
eli
α(t) = (a1 t, . . . , an t)
where we must choose the domain (−ε, ε) to be sufficiently small such that α(t) still lies
in φ(U ).
Next, we map α to M via φ−1 . Define c := φ−1 ◦ α : (−ε, ε) → M . Then
©F
83
By the chain rule and the fact that charts are tangent space isomorphisms,
d
0 −1
c (0) = (φ )∗ α∗ chain rule
dt 0
!
∂
previous proposition
X
= (φ−1 )∗ ai i
i
∂r 0
ou
∂
tangent space isomorphism
X
= ai i
i
∂x p
= Xp .
we can now interpret the abstract definition of a tangent vector geometrically as a directional
derivative using curves.
Proposition 3.1.15
Zh
Suppose Xp is a tangent vector at a point p ∈ M of a manifold and f ∈ Cp∞ (M ). If
c : (−ε, ε) → M is a smooth curve starting at p with c0 (0) = Xp , then
d
Xp f = (f ◦ c).
dt 0
Proof
By the definitions of c0 (0) and c∗ ,
Xp f = c0 (0)f
x
d
= c∗ f
dt 0
d
eli
= (f ◦ c).
dt 0
We have two ways of computing the differential of a smooth map: the function definition
©F
Proposition 3.1.16
Let F : N → M be a smooth map between manifolds, p ∈ N , and Xp ∈ Tp N . If c is a
smooth curve starting at p ∈ N with velocity Xp at p, then
84
Thus F∗,p (Xp ) is the velocity vector of the image curve F ◦ c at F (p).
Proof
By assumption, c(0) = p and c0 (0) = Xp . Thus
ou
= (F∗,p ◦ c∗,0 )
dt
0
d
= (F ◦ c)∗,0
dt 0
0
= (F ◦ c) (0).
Zh
Let g ∈ GL(n, R) and `g : GL(n, R) → GL(n, R) denote left multiplication by g. Since
GL(n, R) is an open subset of the vector space Rn×n , the tangent space Tg (GL(n, R)) can
be identified with Rn×n by mapping partial derivatives to canonical basis vectors. Let
Φ : Tg (GL(n, R)) → Rn×n denote this identification.
Let X ∈ TId (GL(n, R)) = Rn×n . To compute (`g )∗,Id (X), we can choose a curve c(t) ∈
GL(n, R) such that c(0) = Id and c0 (0) = X. From a previous proposition, this means
that c0 (0)f = (f ◦˙ c)(0) for any f ∈ CId
∞
(GL(n, R)). We have shown that such a curve
always exists.
Then `g (c(t)) = gc(t) is simply matrix multiplication and
x
(`g )∗,Id (X) = (`g ◦ c)0 (0) previous proposition
Φ((`g )∗,Id (X)) = Φ((`g ◦ c)0 (0))
= (`g ◦ ˙c)(0) previous proposition
eli
= g ċ(0)
= gΦ(X)
85
Definition 3.1.7 (Immersion)
A smooth map F : N → M between manifolds is said to be an immersion at p ∈ N
if its differential F∗,p : Tp N → TF (p) M is injective. If this holds at all p ∈ N , we say
F is an immersion.
ou
A smooth map F : N → M between manifolds is said to be an submersion at p ∈ N
if its differential F∗,p : Tp N → TF (p) M is surjective. If this holds at all p ∈ N , we say
F is an submersion.
Remark 3.1.18 Recall that if N, M have dimensions n, m respectively, then dim Tp N and
dim TF (p) M = m. The injectivity of the differential F∗,p implies immediately that n ≤ m.
Similarly, surjectively implies n ≥ m.
Zh
Example 3.1.19
The prototype of an immersion is the inclusion of Rn in a higher-dimensional Rm :
Recall the rank of a linear transformation between finite-dimensional vector spaces is the
dimension of the image.
©F
86
Note that the differential is independent of coordinate charts, and so is the rank of the
Jacobian matrix.
ou
Thus p is a regular point of F if and only if F is a submersion at p.
Note we do not define regular values as the image of some regular point. We require all
Zh
points in the pre-image to be regular. On the other hand, c ∈ M is critical if there is a single
critical point in the pre-image.
Proposition 3.1.21
Let f : M → R be smooth. A point p ∈ M is critical if and only if relative to some
chart (U, x1 , . . . , xn ) containing p, all the partial derivatives satisfy
∂f
(p) = 0
∂xj
for j ∈ [n].
x
Proof
The differential f∗,p : Tp M → Tf (p) R ∼ = R is represented by the Jacobian matrix. Since
eli
the image is a linear subspace of R, it is either the zero map or a surjective map. Thus
f∗,p fails to be surjective if and only if all partial derivatives are zero.
Proposition 3.1.22
Let M, N be manifolds and πM : M × N → M, πN : M × N → N be two projections.
For (p, q) ∈ M × N ,
(πM ∗ , πN ∗ ) : T(p,q) (M × N ) → Tp M × Tq N
is an isomorphism.
87
Proof
Recall that if (U, x1 , . . . , xn ) and (V, y 1 , . . . , y m ) are charts about p ∈ M, q ∈ N respec-
tively, (U × V, πM x1 , . . . , πN y 1 , . . . ) is a chart about (p, q) ∈ M × N . Write x̄i := πM xi
and ȳ i = πN y i . We have
!
∂ ∂
πM ∗ xi = (xi ◦ πM )
ou
j
∂ x̄ (p,q) ∂ x̄ j
∂ x̄i
=
∂ x̄j
= δij .
Hence !
∂ ∂
πM ∗ = .
Zh
∂ x̄j (p,q) ∂xj p
Proposition 3.1.23
Let G be a Lie group equipped with its multiplication map µ : G × G → G and inverse
map ι : G → G. The differential of µ at the identity e ∈ G is addition:
µ∗,(e,e) : Te G × Te G → Te G
(Xe , Ye ) 7→ Xe + Ye
x
while the differential of ι at the identity e ∈ G is negation
ι∗,e : Te G → Te G
eli
Xe 7→ −Xe .
Proof
For the first claim, it suffices to show that µ∗, (e, e)(Xe , 0) = Xe and µ∗,(e,e) (0, Ye ) = Ye .
The result then follows by linearity.
©F
To compute µ∗,(e,e) (Xe , 0), We construct a curve α(t) in G × G at which α(0) = (e, e) and
α0 (0) = (Xe , 0) as follows. Take any curve c(t) in G such that c(0) = e and c0 (0) = Xe .
Then define α(t) := (c(t), e). We have
µ∗,(e,e) (Xe , 0) = (µ ◦ α)0 (0)
= c0 (0)
= Xp
The computation for (0, Ye ) 7→ Ye is identical.
88
For the second claim, we need only show that Xe − ι∗,e (Xe ) = 0. Indeed, take the
same curve c(t) in G as above but define α(t) := (c(t), (ι ◦ c)(t)). Note that α0 (0) =
(c0 (0), (ι ◦ c)0 (0)) by construction, which can be checked by evaluating both sides at the
coordinate functions. Then
e = (µ ◦ α)(t)
0 = (µ ◦ α)0 (t)
ou
= µ∗,(e,e) (Xe , (ι ◦ c)0 (0))
= Xe + (ι ◦ c)0 (0)
= Xe + ι∗,e (Xe ).
Zh
chart (V, x1 , . . . , xn ) about p such that
∂
(Xi )p = .
∂xi p
Proof
Let (U, y 1 , . . . , y n ) be any chart about p. Write
X ∂
(Xj )p = aij .
x
i
∂y i p
Since the (Xj )p ’s are linearly independent, the matrix given by A = [aij ] is non-singular,
and we can define a new coordinate system x1 , . . . , xn by
eli
X
yi = aij xj .
j
In other words, xj = )i y .
−1 j i
P
i (A
By the change of basis formula, for which we recall can be verified by evaluating both
©F
sides at y k ,
∂ X ∂y i ∂
=
∂xj i
∂xj ∂y i
X ∂
= aij i .
i
∂y
89
Definition 3.1.12 (Local Maxima)
A real-valued function f : M → R on a manifold is said to have a local maximum at
p ∈ M if there is a neighborhood U 3 p such that f (p) ≥ f (q) for all q ∈ U .
ou
derivative in terms of Newton quotients.
Proposition 3.1.25
A local maximum of a smooth function f : M → R is a critical point of f .
Proof
We need to show that f∗,p is not surjective, meaning it is the zero map, or sends any
Zh
tangent vector to 0.
Fix Xp ∈ Tp M . Let c(t) be a curve starting at p with initial velocity Xp . then f ◦ c is a
differentiable function on an interval with a local maximum at t = 0 so that (f ◦c)0 (0) = 0.
But then
f∗,p (Xp ) = (f ◦ c)0 (0) = 0
for any Xp ∈ Tp M , which concludes the proof.
3.2 Submanifolds
x
Currently, we can check that a given topological space is a manifold either by definition or
by exhibiting it as an appropriate quotient space. We now derive another way: exhibiting
the topological space as a (regular) submanifold of another manifold.
eli
3.2.1 Submanifolds
90
On U ∩ S, φ = (x1 , . . . , xk , 0, . . . , 0). Let φS : U ∩ S → Rk be the restriction of the first k
components of φ to U ∩ S, so φS = (x1 , . . . , xk ). We call such a chart (U, φ) in N an adapted
chart relative to S. Note that (U ∩ S, φS ) is a chart for S in the subspace topology.
ou
Remark 3.2.1 We remark that as a topological space, a regular submanifold of N is re-
quired to have the subspace topology.
Zh
The dimension k of the submanifold may be equal to n, the dimension of the manifold.
In this case, the charts of the submanifold have domain U ∩ S = U . Thus an open subset
of a manifold is a regular submanifold of the same dimension.
Example 3.2.3
The interval S = (−1, 1) on the x-axis is a regualr submanifold of the xy-plane. We can
take the open square (−1, 1) × (−1, 1) as an adapted chart. Then U ∩ S is precisely the
zero set of y on U .
91
Proof
Let (U, φ) = (U, x1 , . . . , xn ) and (V, Ψ) = (V, y 1 , . . . , y n ) be two adapted charts in the
given collection. Assume that they intersect. As we remarked in the definition of a
regular submanifold, we can renumber of coordinates of any adapted chart relative to
a submanifold S so that the last n − k coordinates vanish on points of S. Then for
p ∈ U ∩ V ∩ S,
(ΨS ◦ φ−1 1 k 1 k
S )(x , . . . , x ) = (y , . . . , y ).
ou
Since Ψ ◦ φ−1 is smooth, the y i ’s are smooth functions of the xj ’s and ΨS ◦ φ−1
S is therefore
smooth as well. Similarly, φS ◦ ΨS is also smooth.
−1
This shows that any two charts in {(U ∩ S), φS } are smoothly compatible. The collection
{(U ∩ S, φS )} is thus a smooth structure on S since the U ∩ S’s cover S by assumption.
Zh
3.2.2 Level Sets of a Function
for some c ∈ M .
The value c ∈ M is called the level of F −1 (c). In the special case of F : N → Rm , we say
x
Z(F ) := F −1 (0) is the zero set of F .
Recall that c is a regular value of F if and only if either c is not in the image of F or at
every p ∈ F −1 (c), the differential F∗,p : Tp N → Tp M is surjective. The inverse image of a
eli
regular value c is called a regular level set. If the zero set of some F : N → Rm is regular, it
is called a regular zero set.
Remark that if a regular level set is nonempty, then the map F : N → M is a submersion
at p. In particular, dim N ≥ dim M .
f (x, y, z) = x1 + y 2 + z 2 − 1.
Now, the only critical point of f is 0, which does not lie on the sphere. Thus all points
on the sphere are regular points of f and 0 is a regular value of f .
92
Let p be a point of S 2 at which (∂f /∂x)(p) = 2x(p) 6= 0. Then the Jacobian matrix of
the map (f, y, z) : R3 → R3 is given by
∂f ∂f ∂f
∂x ∂y ∂z
0 1 0
0 0 1
and the Jacobian determinant is nonzero. Hence the inverse function applies and we can
ou
find a neighborhood Up of p ∈ R3 such that (Up , f, y, z) is a chart in the atlas of R3 . In
this chart, the set Up ∩ S 2 is defined by the vanishing of the first coordinate f . Thus
(Up , f, y, z) is an adapted chart relative to S 2 and (Up ∩ S 2 , y, z) is a chart for S 2 .
Similarly, if either (∂f /∂y)(p) 6= 0 or (∂f /∂z)(p) 6= 0, then we can find an adapted chart
(Vp , x, f, z) or (Vp , x, y, f ) containing p in which the set Vp ∩S 2 is the zero set of the second
or third coordinate f . But at least one of the partial derivatives must be nonzero, so as p
Zh
varies over all points of the sphere, we obtain a collection of adapted charts of R3 covering
S 2 . Hence S 2 is a regular submanifold of R3 with dimension 2.
This is an important example since we can nearly translate it verbatim for the regular zero
set of a function F : N → R. First, we note that any regualr level set g −1 (c) of a smooth
function g on a manifold can be expressed as a regular zero set. Indeed, consider f = g − c
so that
g(p) = c ⇐⇒ f (p) = 0.
Moreover, the differentials of f, g are point-wise equal and so f, g have the exact same critical
points. So if g −1 (c) is a regular level set, then f −1 (0) is a regular zero set.
x
Theorem 3.2.7
Let g : N → R be a smooth function on the manifold N . Then a non-empty regular
level set S = g −1 (c) is a regular submanifold of N with codimension 1.
eli
Proof
Let f = g − c so that S = f −1 (0) is the regular zero set of f . Let p ∈ S. Since p is a
regular point of f , relative to any chart (U, x1 , . . . , xn ) about p, (∂f /∂xi )(p) 6= 0 for some
i. By relabelling, we can assume without loss of generality that i = 1. The Jacobian of
the smooth map (f, x2 , . . . , xn ) : U → Rn is given by
©F
∂f
∂x 1 . . . . . . . . .
0 1 ... 0
.. .. . . .. .
. . . .
0 0 ... 1
This is nonzero at the point p by construction. Thus the inverse function theorem applies
and there is a neighborhood Up of p on which f, x1 , . . . , xn forms a coordinate system.
93
Relative to the chart (Up , f, x2 , . . . , xn ), the level set Up ∩ S is defined by setting the first
coordiante f to 0, hence the chart is adapted relative to S. But p ∈ N was arbitrary and
so S is a regular submanifold of codimension 1 in N .
ou
3.2.3 The Regular Level Set Theorem
Our next step is to extend the previous hteorem to a regular level set of a map between
smooth manifolds. This useful theorem is known under various names such as the implicit
function theorem, preimage theorem, and the regular level set theorem. We will follow the
Zh
last.
Proof
Choose a chart (V, Ψ) = (V, y 1 , . . . , y m ) of M centered at c such that Ψ(c) = 0 ∈ Rm .
Then F −1 (V ) is an open set in N containing F −1 (c). Moreover, in F −1 (V ), F −1 (c) =
x
(Ψ ◦ F )−1 (0) and the level set F −1 (c) is the zero set of Ψ ◦ F . If F i := y i ◦ F = ri ◦ (Ψ ◦ F ),
then F −1 (c) is also the common zero set of the functions F 1 , . . . , F m on F −1 (V ).
Since we assumed the regular level set to be nonempty, we must have n ≥ m. Fix a point
eli
p ∈ F −1 (c) and let (U, x1 , . . . , xn ) be a coordinate neighborhood of p ∈ N contained in
F −1 (V ). Since F −1 (c) is a regular level set, p ∈ F −1 (c) is by definition a regular point
of F . Thus the m × n Jacobian matrix of F has rank m. By relabelling if necessary, we
may assume without loss of generality that the first m × m block is nonsingular.
Replace the first m coordinates x1 , . . . , xm of the chart (U, φ) by F 1 , . . . , F m . We claim
that there is a neighborhood Up of p such that (Up , F 1 , . . . , F m , xm+1 , . . . , xn ) is a chart in
©F
the atlas of N . It suffices to compute the Jacobian matrix of the chart function at p. But
the first m × m block is nonsingular by construction and hence the the inverse function
theorem applies.
In the chart (Up , F 1 , . . . , F m , xm+1 , . . . xn ), the set S := F −1 (c) is obtained by setting the
first m coordinates to 0. Since this is true for every point p ∈ S, S is by definition a
regular submanifold of N with codimension m.
The proof of the regular level set theorem yields the following lemma.
94
Lemma 3.2.9
Let F : N → Rm be a smooth map on a manifold N of dimension n and let S be
the level set F −1 (0). Suppose relative to some coordinate chart (U, x1 , . . . , xn ) about
p ∈ S, the determinant of the Jacobian matrix with respect to xj1 , . . . , xjm is nonzero.
Then in some neighborhood of p, we can replace xj1 , . . . , xjm by F1 , . . . , Fm to obtain
an adapted chart for N relative to S.
ou
Remark 3.2.10 The regular level set theorem gives a sufficient but not necessary condition
for a level set to be a regular manifold. For example, if f : R2 → R is the map f (x, y) = y 2 ,
then the zero set Z(f ) = Z(y 2 ) is the x-axis, a regular submanifold of R2 . But ∂f /∂x =
∂f /∂y = 0 on the x-axis, and every point in Z(f ) is a critical point of f . Thus, although
Z(f ) is a regular submanifold of R2 , it is not a regular level set of f .
Zh
3.2.4 Examples of Regular Manifolds
Example 3.2.12
The subset S ⊆ R3 satisfying
x
x3 + y 3 + z 3 = 1
x+y+z =0
is a 1-manifold.
eli
This can be checked by consider the function
F (x, y, z) = (x3 + y 3 + z 3 , x + y + z)
and checking that its Jacobian has rank 2 for every point in S. This implies S is a regular
level set of F and is thus a manifold.
©F
95
(i, j)-minor of A. Recall the cofactor expansion formula
∂f
= (−1)i+j mij .
ou
∂aij
3.2.5 Transversality
p ∈ f −1 (S),
Zh
Definition 3.2.4 (Transversal)
A C ∞ map f : N → M is said to be transversal to a submanifold S ⊆ M if for every
f∗ (Tp N ) = Tf (p) M
for every p ∈ f −1 (c), ie f∗,p is surjective at every p ∈ f −1 (c), ie f −1 (c) is a regular level set.
Thus the transversality theorem is a generalization of the regular level set theorem. It is
©F
useful in giving conditions under which the intersection of two submanifolds is a submanifold.
Proof
Let p ∈ f −1 (S) and (U, x1 , . . . , xm ) be an adapted chart centered at f (p) for M relative
to S such that U ∩ S = Z(xm−k+1 , . . . , xm ), the zero set of the functions xm−k+1 , . . . , xm .
Define g : U → Rk to be the map
g = (xm−k+1 , . . . , xm ).
96
Consider g ◦ f : f −1 (U ) → Rk . By construction,
f −1 (U ) ∩ f −1 (S) = f −1 (U ∩ S)
= {p ∈ f −1 (U ) : g(f (p)) = 0}
= (g ◦ f )−1 (0).
ou
to show that (g ◦ f )∗,p is surjective at every p ∈ f −1 (U ∩ S). Fix Z ∈ T0 Rk . g∗ is by
construction surjective at every point in U as g is a subset of coordinate functions. Hence
there is some Y ∈ Tf (p) M such that g∗ (Y ) = Z. the transversality of f implies that there
is some X ∈ Tp N and Y 0 ∈ Tf (p) S such that f∗ (X) + Y 0 = Y . Since g is constant on S
and f (p) ∈ U ∩ S, we must have g∗ (Y 0 ) = 0 so that
Z = g∗ (Y )
Zh
g∗ (f∗ (X) + Y 0 )
= (g ◦ f )∗ (X) + 0
= (g ◦ f )∗ (X).
But
eli
(g ◦ f )−1 k −1 −1
∗ (T0 R ) = f∗ (g∗ (T0 R ))
k
Corollary 3.2.15.1
If S 0 ⊆ M is a regular submanifold that intersects the regular submanifold S transversely,
then S ∩ S 0 is a regular submanifold of M whose codimension is equal to the sum of the
codimensions of S, S 0 .
97
Proof
Apply the transversality theorem with f = ιS 0 as the inclusion map S 0 → M . Then
f −1 (S) = S ∩ S 0 is a regular submanifold of S 0 . The codimension can be seen from the
definition of a regular submanifold.
ou
some regular submanifold Q ⊆ N . Then for almost every s ∈ S, the map Fs : M → N
given y Fs (x) = F (x, s) is transverse to Q as well.
The proof of this theorem requires an application of Sard’s theorem, which we will see after.
Proof
Since F t Q, W := F −1 (Q) is a regular submanifold of M × S by the transversality
Zh
theorem. Consider the projection on the second factor, π : M × S → S given by π(x, s) =
s. By Sard’s theorem, it suffices to show that whenever s ∈ S is a regular value of
π|W : W → S, then FS t Q.
Fix a regular value s ∈ S of π|W and consider any y ∈ Fs−1 (Q) ⊆ M . Write q := Fs (y) ∈ Q.
Since (y, s) ∈ (π|W )−1 (Q) by construction, (π|W )∗,(y,s) : T(y,s) W → Ts S is surjective by
the definition of a regular value. Moreover, F∗ (T(y,s) W ) = TFs (y) Q by the remark above
since F t Q.
Fix any Zq ∈ TFs (y) N . We need to find some Yq ∈ TFs (y) Q, Xy ∈ Ty M such that
x
Zq = Yq + (Fs )∗ (Xy ).
But (π|W )∗ (T(y,s) W ) = Ts S so we can find some (Xy00 , Xs ) ∈ T(y,s) W such that
(π|W )∗ (Xy00 , Xs ) = Xs .
98
3.2.6 Useful Results
ou
Proof
Let p ∈ S and a neighborhood U 3 p of S such that on U ∩ S, there is a smooth function
f : A → B such that V := A × B ⊆ U and y = f (x) for all (x, y) ∈ V ∩ S. Consider the
function F : V → R2 given by
Zh
The Jacobian is given by
1 0
JF (x, y) = ∂f
∂x
1
Thus F is a diffeomorphism onto its image and can be used as a coordinate map. But
in the chart (V, x, y − f (x)), V ∩ S is defined by the vanishing of the second coordinate
y − f (x). S is by definition a regular submanifold of R2 .
Proposition 3.2.18
The graph Γ(f ) of a smooth function f : R2 → R
x
Γ(f ) := {(x, y, f (x, y)) ∈ R3 }
is a submanifold of R3 .
eli
Proof
Let p ∈ Γ(f ). Consider the function F : R3 → R3 given by
1 0 0
JF (x, y, z) = 0 1 0 .
− ∂x − ∂f
∂f
∂y
1
Thus F is a local diffeomorphism and can be used as a coordinate map. Moreover, S is
globally defined by the vanishing of the third coordinate z − f (x, y) and so is a regular
submanifold of R3 .
Recall a homogeneous polynomial of degreeP k F (x0 , . . . , xn ) ∈ R[x0 , . . . , xn ] is a linear com-
bination of monomials x0 . . . xn of degree j ij = k.
i0 in
99
Lemma 3.2.19 (Euler’s Formula)
Let F (x0 , . . . , xn ) be a homogeneous polynomial of degree k. For any t ∈ R,
ou
i
Proof
Differentiate with respect to t.
On a projective space RP n , a homogeneous polynomial F (x0 , . . . , xn ) of degree k is not a
function, since its value at a point is not necessarily unique. However, the zero set in RP n
of a homogeneous polynomial F (x0 , . . . , xn ) is well defined, since F (a0 , . . . , an ) = 0 if and
Zh
only if
F (ta0 , . . . , tan ) = tk F (a0 , . . . , an ) = 0
for all t ∈ R× .
Proposition 3.2.20
The hypersurface Z(F ) defined by F (x0 , x1 , x2 ) = 0 is smooth if the partial derivatives
eli
are not simultaneously zero on Z(F ).
Proof
Let p ∈ Z(F ). We claim that at least one of ∂F/∂x1 , ∂F/∂x2 is non-zero at p. Suppose
otherwise, But then
©F
X ∂F ∂F
0 = kF (p) = xi = x0 .
i
∂xi ∂x0
But x0 6= 0 on U0 so all partial derivatives vanish, a contradiction.
Recall the standard coordinates on U0 ' R2 are x = x1 /x0 , y = x2 /x0 . In U0 ,
Define f (x, y) := F (1, x, y) so that f, F have the same zero set in U0 . Now, the Jacobian
100
of f is given by ∂F
Jf (x, y) = ∂x .
∂F
∂y
But at least one of this non-zero. We can similarly show this for U1 , U2 .
All in all, Z(F ) is a regular level set of F and is hence a regular submanifold of RP n .
ou
Proposition 3.2.21 (Product of Regular Submanifolds)
If Si is a regular submanifold of the manifold Mi , i = 1, 2, then S1 × S2 is a regular
submanifold of M1 × M2 .
Proof
Fix some (p, q) ∈ S1 × S2 as well as adapted charts (U, x1 , . . . , xn ), (V, y 1 , . . . , y m ) relative
to S1 , S2 and about p, q, respectively so that S1 ∩ U, S2 ∩ V are defined by the vanishing
Zh
of the last k, ` coordinates, respectively. Then (U × V, x1 , . . . , xn , y 1 , . . . , y m ) is a chart
about (p, q) in the product manifold M1 × M2 . By construction, (S1 × S2 ) ∩ (U ∩ V ) is
defined by the vanishing the same k + ` coordinates. It follows by definition that S1 × S2
is a regular submanifold of M1 × M2 with codimension k + `.
Recall that the complex special linear group SL(n, C) is the subgroup of GL(n, C) consisting
of complex matrices with determinant 1.
Proposition 3.2.22
SL(n, C) is a regular submanifold of GL(n, C).
x
Proof
It suffices to show that SL(n, C) is a regular level set of det. That it is a level set is clear.
Since det is complex-valued, it suffices to show that the differential at every A ∈ SL(n, C)
eli
is not identically zero.
Let A ∈ SL(n, C) and consider the curve
A(t) := (1 + t)A
which starts at A(0) = A with initial velocity A ∈ TA Cn×n (under the appropriate iden-
©F
101
3.3 Categories and Functors
ou
Zh
Many problems in mathematics share common features. In topology, one is interested in
knowing if two topological spaces are homeomorphic and in groups theory we wish to know
if two groups are isomorphic. This has given rise to the theories of categories and functors.
A category is essentially a collection of objects and arrow (morphisms) between objects.
These arrows satisfy the abstract properties of maps and are often structure preserving. For
instance, smooth manifolds and smooth maps form a category and so do vector spaces and
x
linear maps. A functor from one category to another preserves the identity morphism and
the composition of morphisms. If the target category is simpler than the domain category,
this provides a way to simply problems in the original category. The tangent space con-
eli
struction with the differential of a smooth map is a functor from the category of smooth
manifolds with a distinguished point to the category of vector spaces. The existence of
the tangent space functor shows that if two manifolds are diffeomorphic, then their tangent
spaces at corresponding points must be isomorphic, thereby proving the smooth invariance
of dimension. Invariance of dimension in the continuous category of topological spaces and
continuous maps is more difficult to show since there is no tangent space functor in the
continuous category.
©F
102
3.3.1 Categories
ou
Definition 3.3.1 (Category)
A category is a collection of elements, called objects, and for any two objects A, B, a
set Mor(A, B) of elements, called morphisms from A to B such that given any f ∈
Mor(A, B), g ∈ Mor(B, C), the composite g ◦ f ∈ Mor(A, C) is defined. Furthermore,
the composition of morphisms satisfies two properties:
(i) (Identity Axiom) for each object A, there is an identity morphism IdA ∈
Zh
Mor(A, A) such that for any f ∈ Mor(A, B) and Mor(B, A),
f ◦ IdA = f, IdA ◦g = g.
(ii) (Associative Axiom) for f ∈ Mor(A, B), g ∈ Mor(B, C) and h ∈ Mor(C, D),
h ◦ (g ◦ f ) = (h ◦ g) ◦ f.
Example 3.3.1
x
The collection of groups and group homomorphisms forms a category where the objects
are groups and Mor(A, B) is the set of groups homomorphisms from A to B.
eli
Example 3.3.2
The collection of vector spaces over R and R-linear maps forms a category where objects
are real vector spaces and Mor(V, W ) is the set of linear maps from V to W .
103
Definition 3.3.2 (Object Isomorphism)
Two objects A, B in a category are isomorphic if there are morphisms f : A → B and
g : B → A such that
g ◦ f = IdA , f ◦ g = IdB .
In this case, both f, g ar called isomorphisms.
ou
The usual notation for an isomorphism is '. Thus A ' B can mean a group isomorphism,
a vector space isomorphism, a homeomorphism, or a diffeomorphism, depending on the
category and the context.
3.3.2 Functors
Zh
Definition 3.3.3 ((Covariant) Functor)
A (covariant) functor F from one category C to another category D is a map that
associates to each object A in C an object F(A) in D and to each morphism f : A → B
a morphism F(f ) : F(A) → F (B). such that
(i) F(IdA ) = IdF (A)
x
(ii) F(f ◦ g) = F(f ) ◦ F(g)
Example 3.3.6
eli
The tangent space construction is a functor fro the category of pointed manifolds to the
category of vector spaces. To each pointed manifold (N, p) we associate the tangent
space Tp N and to each smooth map f : (N, p) → (M, f (p)) we associate the differential
f∗,p : Tp N → Tf (p) M .
The functorial property (i) holds because if Id : N → N is the identity map, then its
©F
differential Id∗,p : Tp N → Tp N is also the identity map. The functorial property (ii) holds
because in this context it is simply the chain rule.
Proposition 3.3.7
Let F : C → D be a functor between categories. If f : A → B is an isomorphism in C,
then F(f ) : F(A) → F (B) is an isomorphism in D.
If in condition (ii) of the definition for a covariant functor we reverse the direction of the
arrow for the morphism F(f ), we obtain a contravariant functor.
104
Definition 3.3.4 (Contravariant Functor)
A contravariant functor from from category C to another category D is a map that
associates to each object A in C an object F(A) in D and to each morphism f : A → B
a morphism F(f ) : F(B) → F (A) such that
(i) F(IdA ) = IdF (A)
(ii) F(f ◦ g) = F(g) ◦ F(f )
ou
Example 3.3.8
Smooth functions on a manifold give rise to a contravariant functor that associates to
each manifold M the algebra F(M ) = C ∞ (M ) of C ∞ functions on M and to each smooth
map F : N → M of manifolds the pullback map F(F ) = F ∗ : C ∞ (M ) → C ∞ (N ) given
by
Zh
F ∗ (h) = h ◦ F
for h ∈ C ∞ (M ). It can be checked that the pullback satisfies the two functorial properties.
L∨ (α) = α ◦ L.
Note that the dual of L reverses the direction of the arrow.
Proposition 3.3.9
Suppose V, W, S are real vector spaces.
(i) If IdV : V → V is the identity map, then Id∨V : V ∨ → V ∨ is the identity map on
V ∨.
(ii) If f : V → W and g : W → S are linear maps, then (g ◦ f )∨ = f ∨ ◦ g ∨ .
105
This proposition shows that the dual construction F : () 7→ ()∨ is a contravariant functor
from the category of vector spaces to itself: for a real vector space V , F(V ) = V ∨ and
for f ∈ Hom(V, W ), F(f ) = f ∨ ∈ Hom(W ∨ , V ∨ ). Consequently, if f : V → W is an
isomorphism, then so is its dual f ∨ : W ∨ → V ∨ .
Fix a positive integer k ≥ 1. For any linear map L : V → W of vector spaces define the
pullback map L∗ : Ak (W ) → Ak (V ) that sends f ∈ Ak (W ) to
ou
(L∗ f )(v1 , . . . , vk ) = f (L(v1 ), . . . , L(vk )).
It can be checked from the definition that L∗ is a linear map.
Proposition 3.3.10
The pullback of covectors by a linear map satisfies the two functorial properties:
(i) If IdV : V → V is the identity map on V , then Id∗V = IdAk (V ) , the identity map on
Ak (V ).
Zh
(ii) If K : U → V and L : V → W are linear maps of vector spaces, then (L ◦ K)∗ =
K ∗ ◦ L∗ : Ak (W ) → Ak (U ).
To each vector space V , we associate the vector space Ak (V ) of all k-covectors on V , and
to each linear map L : V → W of vector spaces, we associate the pullback Ak (L) = L∗ :
Ak (W ) → Ak (V ). Then Ak () is a contravariant functor from the category of vector spaces
and linear maps to itself.
When k = 1, the space A1 (V ) is simply the dual space, and for any linear map L : V → W ,
the pullback map A1 (L) = L∗ is the dual map L∨ : W ∨ → V ∨ . Thus the multicovector
x
functor Ak () generalizes the dual functor ()∨ .
106
The image of a smooth map does not in general have a nice structure. However, using
the immersion theorem we derive conditions under which the image of a smooth map is a
manifold.
ou
Suppose f : N → M is a smooth map of manifolds and we want to show that the level
set f −1 (c) is a manifold for some c ∈ M . A sufficient condition from the regular level set
theorem is for the differential to have maximal rank at every point of f −1 (c). However, even
if this is true, it may be difficult to show at times. The constant rank has the comparative
advantage in that it is not necessary to know the precise rank of f ; it suffices that the rank
be constant.
Let us recall the constant rank theorem for Euclidean spaces.
Zh
Theorem 3.4.1 (Constant Rank in Euclidean Space)
If f : U ⊆ Rn → Rm has constant rank k in a neighborhood of a point p ∈ U , there
are diffeomorphisms G of a neighborhood of p ∈ U sending p 7→ 0 ∈ Rn and F of a
neighborhood of f (p) ∈ Rm sending f 7→ 0 ∈ Rm such that
Thus after a suitable change of coordinates near p ∈ U and f (p) ∈ Rm , the map f
assumes the form
x
(x1 , . . . , xn ) 7→ (x1 , . . . , xk , 0, . . . , 0).
Proof
eli
By reordering the f 1 , . . . , f m and coordinates x1 , . . . , xn if necessary, we may assume
without loss of generality that the first k × k-submatrix of the Jacobian JF (p) at the
point p is non-singular. Relabel the coordinates (x, y) = (x1 , . . . , xk , y 1 , . . . , y n−k ) := x
and the function (f, g) = (f 1 , . . . , f k , g 1 , . . . , g n−k ) := f .
Define G : U → Rm by
©F
107
On V1 ,
(u, v) = (G ◦ G−1 )(u, v) = (f ◦ G−1 , y ◦ G−1 )(u, v).
Comparing the first components give u = (f ◦ G−1 )(u, v). Hence
ou
= (u, h(u, v)). h := g ◦ G−1
Zh
For this matrix to have constant rank k, ∂g/∂v must be identically zero on V1 . Thus h is
a function of u alone and we can write
We have
x
(F ◦ f ◦ G−1 )(u, v) = F (u, h(u))
= (u, h(u) − h(u))
eli
= (u, 0).
The constan rank theorem for Euclidean spaces has an immediate analogue for manifolds.
Proof
Choose a chart (Ū , φ̄) about p ∈ N and (V̄ , Ψ̄) about f (p) ∈ M . Then Ψ̄◦f ◦ φ̄−1 is a map
between open subsets of Euclidean spaces. Because φ̄, Ψ̄ are diffeomorphisms, Ψ̄ ◦ f ◦ φ̄−1
108
has the same constant rank k as f in a neighborhood of φ̄(p) ∈ Rn . By the constant
rank theorem for Euclidean spaces, there are diffeomorphisms G of a neighborhood of
φ̄(p) ∈ Rn and F of a neighborhood of (Ψ̄ ◦ f )(p) ∈ Rm such that
ou
The constant-rank level theorem follows easily. Recall a neighborhood of a subset A ⊆ M is
an open set containing A.
Zh
Proof
Let p ∈ f −1 (c) be arbitrary. By the constant rank theorem, we can find a coordinate chart
(U, φ) = (U, x1 , . . . , xn ) centered at p ∈ N and a coordinate chart (V, Ψ) = V (y 1 , . . . , y m )
centered at f (p) = c ∈ M such that
This shows that the level set (Ψ◦f ◦φ−1 )−1 (0) is defined by the vanishing of the coordinates
r1 , . . . , rk . The image of the level set f −1 (c) under φ is precisely the level set (Ψ ◦ f ◦
x
φ−1 )−1 (0), since
f (A) := AT A.
O(n) = f −1 (I).
It can be checked that f is of constant rank on GL(n, R) and hence O(n) is a regular
submanifold of GL(n, R).
109
Consider f : N → M a map with constant rank k in a neighborhood of a point p ∈ N ,
with charts (U, φ) = (U, x1 , . . . , xn ) about p and (V, Ψ) = (V, y 1 , . . . , y m ) about f (p) from
the constant rank theorem. Note that for any q ∈ Q,
ou
= (Ψ ◦ f ◦ φ−1 )(φ(q))
= (Ψ ◦ f ◦ φ−1 )(x1 (q), . . . , xn (q))
= (x1 (q), . . . , xk (q), 0, . . . , 0).
As functions on U ,
(y 1 ◦ f, . . . , y m ◦ f ) = (x1 , . . . , xk , 0, . . . , 0).
Zh
The local normal form of f relative to the charts above in the constant rank theorem can be
expressed in terms of the local coordinates x1 , . . . , xn and y 1 , . . . , y m as follows. The map f
is given by
(x1 , . . . , xn ) 7→ (x1 , . . . , xk , 0, . . . , 0).
The constant rank theorem givrs local normal forms for immersions and submersions, called
the immersion theorem and submersion theorem respectively.
x
Consider a smooth map f : N → M between manifolds of dimension n, m respectively. Let
(U, (xi )), (V, (y j )) be charts about p ∈ N, f (p) ∈ M respectively and consider the Jacobian
of f∗,p with respect to these charts. Then for any p ∈ N , f∗,p is injective if and only if n ≤ m
eli
and the Jacobian has rank rank J = n. Similarly, f∗,p is surjective if and only if n ≥ m and
rank J = m.
Having maximal rank at a point is an open condition in the sense that the set
is equivalent to the vanishing of all k × k minors of the Jacobian. As the pullback of a closed
(singleton) set of finitely many continuous functions, U − Dmax (f ) is closed and so Dmax (f )
is open. In particular, if f has maximal rank at p, then it has maximal rank at all points in
some neighborhood of p.
110
Proposition 3.4.5
Let N be an n-manifold and M be an m-manifold. If a smooth map f : N → M is an
immersion at a point p ∈ N , then it has constant rank n in a neighborhood of p. If a
smooth map f : N → M is a submersion at a point p ∈ N , then it has constant rank m
in a neighborhood of p.
The following theorems follow immediately as special cases of the constant rank theorem.
ou
Theorem 3.4.6 (Immersion/Submersion)
(i) (Immerision theorem) Suppose f : N → M is an immersion at p ∈ N . Then
there are charts (U, φ) centered at p ∈ N and (V, Ψ) centered at f (p) ∈ M such
that in a neighborhood of φ(p),
Zh
(ii) (Submerision theorem) Suppose f : N → M is an submersion at p ∈ N . Then
there are charts (U, φ) centered at p ∈ N and (V, Ψ) centered at f (p) ∈ M such
that in a neighborhood of φ(p),
Corollary 3.4.6.1
A submersion f : N → M of manifolds is an open map.
x
Proof
Let W ⊆ N be open and pick a point f (p) ∈ f (W ). By the submersion theorem, there are
charts (U, φ), (V, Ψ) about p, f (p) such that (Ψ ◦ f ◦ φ−1 ) is a projection. Let U ⊇ B 3 p
eli
be an open neighborhood about p. Then φ(B) is open in Rn . But projections are open
maps and so
(Ψ ◦ f )(B) = (Ψ ◦ f ◦ φ−1 )(φ(B))
is open in Rm . But then
f (B) = Ψ−1 [(Ψ ◦ f )(B)]
©F
is an open subset of f (W ) containing f (p). This concludes the proof by the arbitrary
choice of f (p).
We now derive the regular level set theorem as corollaries of both the submersion and constant
rank theorems. Indeed, for a smooth map f : N → M of manifolds, a level set f −1 (c) is
regular if and only if f is a submersion at every point p ∈ f −1 (c). Fix one such point
p ∈ f −1 (c) and let (U, φ), (V, Ψ) be charts in the submersion theorem. Then Ψ ◦ f ◦ φ−1 =
111
π : φ(U ) → Rm is the projection onto the first m coordinates,
π(r1 , . . . , rn ) = (r1 , . . . , rm ).
It follows that on U ,
Ψ ◦ f = π ◦ φ = (r1 , . . . , rm ) ◦ φ = (x1 , . . . , xm ).
ou
It follows that
Zh
A third proof of the regular level set theorem using the submersion theorem proceeds as
follows. On a regular level set f −1 (c), the map f : N → M has maximal rank m at every
point. Since the maximality of the rank is an open condition, a regular level set f −1 (c) has
a neighborhood on which f has constant rank m. By the constant rank level set theorem
above, f −1 (c) is thus a regular submanifold of N .
Example 3.4.7
f (t) = (t2 , t3 ) is not an immersion since f 0 (0) = (0, 0).
eli
Example 3.4.8
f (t) = (t2 − 1, t3 − t) is an immersion since the equation
has no solutions in t.
Example 3.4.9
Let M be the union of the graph of y = sin(1/x) on (0, 1) and a smooth curve joining (0, 0)
and (1, sin 1). f : R → M in the intuitive way is a injective immersion whose image with
the subspace topology is not homeomorphic to R.
In the examples above, f (N ) is not a regular submanifold of M = R2 . We would like
conditions on f so that its image would be a regular submanifold of M .
112
Definition 3.4.1 (Embedding)
A C ∞ map f : N → M is called an embedding if
(i) it is an (injective) immersion
(ii) the image f (N ) with the subspace topology is homeomorphic to N under f .
Note that the condition that f is injective in (i) is redundant since a homeomorphism is
ou
necessarily a bijection.
Remark 3.4.10 The word “submanifold” is used differently in many contexts. Some au-
thors give the image f (N ) of an injective immersion the topology inherited from f rather
than the subspace topology of M . With this topology, f (N ) is by definition homeomorphic
to N . These authors define a submanifold to be the image of any injective immersion with
the topology and differentiable structure inherited from f . Such a set is sometimes called an
immersed submanifold. Note that if the underlying set of an immersed submanifold is given
Zh
the subspace topology, it need not be a manifold at all!
For us, a submanifold without any qualifying adjective is always a regular submanifold.
We use the phrase “near p” to mean “in a neighborhood of p.”
Theorem 3.4.11
If f : N → M is an embedding, then its image f (N ) is a regular submanifold of M .
Proof
x
Fix p ∈ N . We need to show that im some neighborhood of f (p), the set f (N ) is defined
by the vanishing of m − n coordinates.
By the immersion theorem, there are local coordinates (U, x1 , . . . , xn ) near p and (V, y 1 , . . . , y m )
eli
near f (p) such that f : U → V has the form
(V ∩ V 0 ) ∩ f (N ) = V ∩ f (U ) = (V ∩ V 0 ) ∩ f (U )
113
Theorem 3.4.12
If N is a regular submanifold of M , then the inclusion ι : N → M is an embedding.
Proof
Since a regular submanifold has the subspace topology and ι(N ) also has the subspace
topology, ι : N → ι(N ) is certainly a homeomorphism. It remains to show that ι : N → M
ou
is an immersion.
Fix p ∈ N . Choose an adapted chart (V, y 1 , . . . , y n , y n+1 , . . . , y m ) for M about p such that
V ∩ N is the zero set of y n+1 , . . . , y m . Relative to the charts (V ∩ N, y 1 , . . . , y n ) for N and
(V, y 1 , . . . , y m ) for M , the inclusion i is given by
(y 1 , . . . , y n ) 7→ (y 1 , . . . , y n , 0, . . . , 0)
Zh
which shows that ι is an immersion.
The image of an embedding is often called an embedded submanifold. Our results above show
that an embedded submanifold and a regular submanifold are the same thing.
Proof
Denote the dimensions of N, M, S by n, m, s, respectively. Fix p ∈ N . Since S is a
regular submanifold of M , there is an adapted coordinate chart (V, Ψ) = (V, y 1 , . . . , y m )
for M about p such that S ∩ V is the zero set of y s+1 , . . . , y m , with coordinate chart
©F
114
particular at p. We conclude the proof by the arbitrary choice of p ∈ N .
given by (A, B) 7→ (A, B) is smooth since each entry of the resulting matrix is a polynomial
ou
function of the entries of the input matrices. However, it is not immediately obvious that
the induced map
µ̄ : SL(n, R) × SL(n, R) → SL(n, R)
is smooth as the canonical coordinate system on GL(n, R) is not a coordinate system on
SL(n, R).
Now, SL(n, R) × SL(n, R) is a regular submanifold of the product manifold GL(n, R) ×
Zh
GL(n, R). Thus the inclusion map
is smooth. It follows thaat the composition µ ◦ ι is smooth as well. Because the image
of µ ◦ ι lies in SL(n, R), a regular submanifold of GL(n, R), we can apply the previous
theorem to deduce that the induced multiplication map is smooth.
3
d X ∂f
0 = f (c(t)) = i
(c(t))(ci )0 (t).
dt i=1
∂x
At t = 0,
3
d X ∂f
0 = f (c(0)) = i
(p)v i .
dt i=1
∂x
115
Since the vector v = hv 1 , v 2 , v 3 i represents the arrow from p = (p1 , p2 , p3 ) to x = (x1 , x2 x3 ) ∈
Tp N , we usually make the substitution v i = xi − pi . This amounts to translating the tangent
plane from the origin to “p”. Thus the tangent plane to N at p is defined by the equation
3
d X ∂f
0 = f (c(0)) = i
(p)(xi − pi ).
dt ∂x
ou
i=1
Zh
of f at p is normal to any vector in the tangent plane.
We remark that to recover the tangent vector v ∈ Tp N , we do not use the substitution and
instead directly compute the v i ’s.
∂f
x
= 2x
∂x
∂f
= 2y
∂y
eli
∂f
= 2z.
∂z
At p = (a, b, c),
∂f
(p) = 2a
∂x
©F
∂f
(p) = 2b
∂y
∂f
(p) = 2c.
∂z
The equation determining the tangent sphere that we derived above is given by
116
3.4.6 The Differential of an Inclusion Map
Let ι : S 1 → R2 be the inclusion map of the unit circle. Denote by x, y the standard
coordinates on R2 and x̄ = i∗ x, ȳ = i∗ y their restrictions to S 1 . On the upper semicircle
ou
x̄ is a local coordinate, so that ∂/∂ x̄ is defined.
We know that
!
∂ ∂ ∂ ∂
u +v = ι∗ = (· ◦ ι)
∂x p ∂y p ∂ x̄ p ∂ x̄ p
Zh
for some u, v ∈ R. Evaluating at x̄, ȳ yields the result
!
∂ ∂ ∂ ȳ ∂
ι∗ = + · .
∂ x̄ p ∂x p ∂ x̄ ∂y p
!
∂ ∂ ∂ ∂
ι∗ = α1 + β1 + γ1
∂u p ∂x p ∂y p ∂z p
!
∂ ∂ ∂ ∂
ι∗ = α2 + β2 + γ2 .
∂v p ∂x p ∂y p ∂z p
117
the point p = (a, b, c),
!
∂
γ 1 = ι∗ (z)
∂u p
∂
= (z ◦ i)
∂u p
√
ou
∂
= ( 1 − u2 − v 2 )
∂u p
−2u
= √
2 1 − u2 − v 2 p
a
=− .
c
2
γ = ...
Zh
b
=− .
c
Proposition 3.4.17
Every smooth map f from a compact manifold N → Rm has a critical point.
Proof
x
Suppose towards a contradiction that f∗,p is surjective at every p ∈ N , ie it is a submersion.
Let π : Rm → R denote the projection onto the first coordinate and consider π◦f : N → R.
It is clear that π is a submersion. But then π ◦ f is a submersion as well.
eli
Recall that continuous maps attain their extrema on compact sets. Such an extreme is
necessarily a critical point of π ◦ f , which contradicts that π ◦ f is a submersion.
Proposition 3.4.18
Any injective immersion f : N → M on a compact manifold N is an embedding.
©F
Proof
An injective map is bijective onto its image f (N ) ⊆ M . Moreover, f is an immersion
by assumption. It suffices to check that f −1 is continuous in order to conclude that f is
homeomorphic onto its image and thus is an embedding.
We check that f is a closed map, so the pre-images of closed sets under f −1 are closed and
so f −1 is continuous. This is purely a topological results. Any closed subset of a compact
space N is compact. Hence the continuous image under f is also compact and therefore
closed.
118
This concludes the proof.
A smooth vector bundle over a smooth manifold M is a smoothly varying family of vector
spaces, parametrized by points of M , that locally looks like a product. The collection of
ou
tangent spaces to a manifold has the structure of a vector bundle over the manifold, called the
tangent bundle. A smooth map between two manifolds induces a bundle map between two
manifolds. This the tangent bundle construction is a functor from the category of smooth
manifolds to the category of vector bundles.
For us, the importance of the vector bundle point of view comes from its role in unifying
concepts. A section of a vector bundle π : M → E is a mapping from each point of M into
the fiber of the bundle over the point. Both vector fields and differential forms on a manifold
Zh
are sections of vector bundles over the manifold.
Let M be a smooth manifold. Recall that at each p ∈ M , the tangent space Tp M is the
vector space of all point-derivations of Cp∞ (M ), the algebra of germs of C ∞ functions at p.
In general, if {Ai }i∈I is an indexed collection of subsets of a set S, then their disjoint union
is defined to be the set G [
Ai := {i} × Ai .
i∈I i∈I
Since the union ∪p∈M Tp is already disjoint, it is not so important to us whether to specify
©F
∪, t in the definition.
There is a natural map π : T M → M given by π(v) = p where v ∈ Tp M . At times, we use
the notation (p, v) to denote a tangent vector v ∈ Tp M to make explicit the point p ∈ M at
which v is a tangent vector.
As defined, T M has no topology or manifold structure. We will make it into a smooth
manifold and further show that it is a smooth vector bundle over M . We focus for now on
defining a topology on T M .
119
Let (U, φ) = (U, x1 , . . . , xn ) be a coordinate chart on M . Define
[ [
TU = Tp U = Tp M.
p∈U p∈U
Recall that a basis for Tp U = Tp M is the set of partial derivatives of coordinate functions.
Thus any v ∈ Tp M can be uniquely written as
ou
X ∂
v= ci .
i
∂xi p
The coeficients ci = ci (v) depend on v and so are functions on T U . Let x̄i = xi ◦ π where
π(v) = p for v ∈ Tp M and define the map φ̃ : T U → φ(U ) × Rn by
v 7→ (x1 (p), . . . , xn (p), c1 (v), . . . , cn (v)) = (x̄1 , . . . , x̄n , c1 , . . . , cn )(v).
Then φ̃ has an inverse
Zh
X ∂
(φ(p), c1 , . . . , cn ) 7→ ci .
i
∂xi p
This means we can use φ̃ to transfer the topology of φ(U ) × Rn to T U : a set in T U is open if
and only if φ̃(A) is open in φ(U ) × Rn , under the standard topology of R2n . By construction,
T U is homeomorphic to φ(U ) × Rn . If V ⊆ U is open, then φ(V ) × Rn is an open subset of
φ(U ) × Rn . Hence the relative topology on T V as a subset of T U is the same as the topology
induced fron the bijection φ̃|T V : T V → φ(V ) × Rn .
Let φ∗ : Tp U → Tφ(p) Rn be the differential of the coordinate map φ at p. We may identity
φ∗ (v) with some column vector hc1 , . . . , cn i ∈ Rn where ci ’s are the coefficients of the tangent
x
vector φ∗ (v) with respect to the standard basis of Tφ(p) Rn . Thus another way to describe φ̃
is
φ̃ = (φ ◦ π, φ∗ ).
eli
Let B be the collection of all open subsets of T (Uα ) for all coordinate open sets Uα in M .
That is,
α
©F
Lemma 3.5.1
(i) For any manifold M , the set T M is the union of all A ∈ B.
(ii) Let U, V be coordinate open sets in a manifold M . If A is open in T U and B
is oepn in T V , then A ∩ B is open in T (U ∩ V ).
It follows from this lemma that B forms a basis for some topology on T M . We given the
tangent bundle T M the topology generated by the basis B.
120
Proof
(i) Let {(Uα , φα )} be an atlas for M . Then
[ [
TM = T (Uα ) ⊆ A ⊆ T M.
α A∈B
ou
(ii) Since T (U ∩V ) is a subspace of T U , A∩T (U ∩V ) must be open in T (U ∩V ). Similarly,
B ∩ T (U ∩ V ) is open in T (U ∩ V ). But then
A ∩ B ⊆ T U ∩ T V = T (U ∩ V )
so that A ∩ B is open in T (U ∩ V ).
Lemma 3.5.2
Zh
A topological manifold M has a countable basis consisting of coordinate open sets.
Proof
Let {(Uα , φα )} be an atlas on M and B = {Bi } a countable basis for M . For each
coordinate open set Uα and p ∈ Uα , there is a basic open set Bp,α ∈ B such that
p ∈ Bp,α ⊆ Uα .
Proposition 3.5.4
The tangent bundle T M of a manifold is Hausdorff.
Proof
Let (p, Xp ) 6= (q, Xq ) ∈ T M .
If p 6= q, then since M is hausdorff, we can find open sets U, V ⊆ M separating p, q. By
shrinking U, V if necessary, we may assume that U, V are coordinate open sets. Then
121
T U, T V are disjoint open sets in T M separating (p, Xp ), (q, Xq ).
Suppose now that p = q and Xp 6= Xq . Let (U, φ) be a chart about p. Then T U is
homeomorphic to an open subset of φ(U ) × Rn ⊆ R2n through the map φ̃. But Euclidean
space is Hausdorff so there are open sets Vp , Vq separating φ̃(p, Xp ), φ̃(p, Xq ). Since φ̃ is a
homeomorphism, φ̃−1 (Vp ), φ̃−1 (Vq ) separate (p, Xp ), (p, Xq ) as desired.
ou
3.5.2 The Manifold Structure on the Tangent Bundle
Our goal is now to show that this is indeed a smooth structure on T M . We already have
T M = ∪α T Uα . It remains to check that φ̃α , φ̃β are smoothly compatible on (T Uα ) ∩ (T Uβ ).
Zh
Recall if (U, x1 , . . . , xn ), (V, y 1 , . . . , y n ) are two charts on M , then for any p ∈ U ∩ V there
are two bases singled out for the tangent space Tp M : {∂/∂xj |p }j and {∂/∂y i |p }i . So any
tangent vector v ∈ Tp M has two descriptions:
X ∂ X ∂
v= aj = bi .
j
∂xj p i
∂y i p
φ̃β ◦ φ̃−1 n
α : φα (Uαβ ) × R → φβ (Uαβ ) × R
n
is given by
!
X ∂
(φα (p), a1 , . . . , an ) 7→ p, aj
j
∂xj p
122
But recall we computed
X ∂y i
bi = aj (p)
j
∂xj
X ∂(φβ ◦ φ−1
α )
i
= aj (φα (p)).
j
∂rj
ou
By the definition of a smooth atlas, φβ ◦φ−1
α is C . Therefore φ̃β ◦ φ̃α is C . This completes
∞ −1 ∞
the proof that the tangent bundle T M is a smooth manifold with {(T Uα , φ̃α )} as a smooth
atlas.
Zh
On the tangent bundle T M of a smooth manifold M , the natural projection map π : T M →
M given by
π(p, v) := p
makes T M into a C ∞ vector bundle over M . We make this precise in this section.
Such an open set U in (ii) is called a trivializing open set for E and φ is a trivialization of E
over U . Also note that φ is fiber-preserving with respect to the projection map U × Rr → U .
The collection {(U, φ)} with {U } being an open cover of M , is called a local trivialization
for E, and {U } is a trivializing open cover of M for E.
123
Definition 3.5.4 (Smooth Vector Bundle)
A C ∞ vector bundle of rank r is a triple (E, M, π) consisting of manifolds E, M and
a surjective smooth map π : E → M that is locally trivial of rank r.
The manifold E is called the total space of the vector bundle and M is the base space. By
abuse of language, we say that E is a vector bundle over M .
ou
Let (E, M, π) be a vector bundle of rank r. For any regular submanifold S ⊆ M , the triple
(π −1 S, S, π|π−1 (S) ) is also a smooth vector bundle over S, called the restriction of E to S.
We often write the restriction as E|S instead of π −1 S.
Properly speaking, the tangent bundle of a manifold M is a triple (T M, M, π), and T M is
the total space of the tangent bundle. Here π is the canonical projection as aforementioned.
In common usage, T M is often referred to as the tangent bundle.
Zh
Example 3.5.5 (Product Bundle)
Given a manifold M , let π : M × Rr → M be the projection onto the first factor. Then
M × Rr → M is a vector bundle of rank r, called the product bundle of rank r over M .
The vector space structure on the fiber π −1 (p) is the obvious one.
A local trivialization on M × R is given by the identity map IdM ×R . The infinite cylinder
S 1 × R is the product bundle of rank 1 over the circle.
Let π : E → M be a smooth vector bundle. Suppose (U, Ψ) = (U, x1 , . . . , xn ) is a chart on
M and
x
φ : E|U → U × Rr
φ(e) = (π(e), c1 (e), . . . , cr (e))
eli
is a trivialization of E over U . Then
(Ψ × Id) ◦ φ
= (x1 , . . . , xn , c1 , . . . , cr ) : E|U → U × Rr → Ψ(U ) × Rr
⊆ Rn × Rr
©F
is a diffeomorphism of E|U onto its image and so is a chart on E. We call x1 , . . . , xn the base
coordinates and c1 , . . . , cr the fiber coordinates of the chart
(E|U , (Ψ × Id) ◦ φ)
on E. Note that the fiber coordinates ci depend only on the trivialization φ of the bundle
E|U and not on the trivialization Ψ of the base U .
Let πE : E → M and πF : F → N be two vector bundles, possibly of different ranks.
124
Definition 3.5.5 (Bundle Map)
A bundle map from E to F is a pair of maps (f, f˜) where f : M → N and f˜ : E → F
such that
(i) πF ◦ f˜ = f ◦ πE
(ii) f˜ is linear on each fiber, ie for each p ∈ M , f˜ : Ep → Ff (p) is a linear map of
vector spaces.
ou
The collection of all vector bundles together with bundle maps between them forms a cate-
gory.
Example 3.5.6
A smooth map f : N → M between manifolds induces a bundle map (f, f˜), where
f˜ : T N → T M is given by
Zh
f˜(p, v) = (f (p), f∗ (v)) ∈ {f (p)} × Tf (p) M ⊆ T M
for all v ∈ Tp N . This gives rise to a covariant functor T from the vategory of smooth
manifolds and smooth maps to the cateogry of vector bundles and bundle maps: To each
manifold M , we associate its tangent bundle T M , and to each smooth map f : N → M
between manifolds, we associate the bundle map
T f = (f : N → M, f˜ : T N → T M ).
If E, F are two vector bundles over the same manifold M , then a bundle map from E to F
x
over M is a bundle map in which the base map is the identity IdM . For a fixed manifold M ,
we can also consider the category of all smooth vector bundles over M and smooth vector
bundles over M . In this category, it makes sense to speak of an isomorphism of vector bundles
over M . Any vector bundle over M isomorphic over M to the product bundle M × Rr is
eli
called a trivial bundle.
This condition simply means that for each p ∈ M , s maps p into the fiber Ep above p. We
can visualize a section as a “cross-section” of the bunndle. We say that a section is smooth
if it is smooth as a map from M → E.
125
Definition 3.5.7 (Vector Field)
A vector field X on a manifold M is a function that assigns a tangent vector Xp ∈ Tp M
to each point p ∈ M .
In terms of the tangent bundle, a vector field on M is imply a section of the tangent bundle
π : T M → M and the vector field is smooth if it is smooth as a map from M → T M .
ou
Example 3.5.7
The formula
∂ ∂ −y
X(x,y) = −y +x =
∂x ∂y x
defines a smooth vector field on R2 .
Let s, t : M → E be smooth sections of a smooth vector bundle π : E → M and f ∈ C ∞ (M ).
Zh
We define the sum s + t : M → E and product f s : M → E as follows:
(s + t)(p) := s(p) + t(p) ∈ Ep p∈M
(f s)(p) := f (p)s(p) ∈ Ep . p∈M
Proposition 3.5.8
Let s, t : M → E be smooth sections of a smooth vector bundle π : E → M and
f ∈ C ∞ (M ).
(i) s + t is a smooth section of E
(ii) f s is a smooth section of E
x
Proof
(i) It is clear that s + t is a section of E. To show that it is smooth, fix a point p ∈ M
eli
and let V 3 p be a trivializing open set for E, with smooth trivialization
φ : π −1 (V ) → V × Rr .
Since s, t are smooth, the components ai , bi are smooth on V . Since φ is linear on each
fiber,
(φ ◦ (s + t))(q) = (q, a1 (q) + b1 (q), . . . , ar (q) + br (q)).
This proves that s + t is a smooth map on V and hence at p. Since p is an arbitrary point
of M , the section s + t is smooth on M .
126
(ii) We again use the linearity of φ along each fiber to deduce that
Since smoothness is preserved under multiplication, this function is smooth. We omit the
details as it is similar to (i).
Denote the set of all smooth sections of E by Γ(E). The proposition above shows that
ou
Γ(E) is not only a vector space over R, but also a module over the ring C ∞ (M ). For any
open subset U ⊆ M , one can also consider the vector space Γ(U, E) of smooth sections of
E over U . Then Γ(U, E) is both a vector space over R and a C ∞ (U )-module. Note that
Γ(M, E) = Γ(E). To contrast with sections over a proper subset U , a section over the entire
manifold M is called a global section.
127
Lemma 3.5.11
Let φ : E|U → U × Rr be a trivialization over an open set U of a smooth vector
bundle E → M , and t1 , . . . , tr the smooth frame over U of the trivialization. Then a
section X
s= bi ti
i
ou
of E over U is smooth if and only if its coefficients bi relative to the frame t1 , . . . , tr
are smooth.
Proof
Suppose the section s = i bi ti of E over U is smooth. Then φ ◦ s is smooth. But
P
X
(φ ◦ s)(p) = bi (p)φ(ti (p))
Zh
i
X
= bi (p)(p, ei )
i
!
X
= p, bi (p)ei .
i
Thus bi (p) are simply the fiber coordinates of s(p) relative to the trivialization φ. Since
φ ◦ s is smooth, all the bi ’s must be smooth.
The converse holds since the collections of smooth sections is a module over C ∞ (U ).
x
Proposition 3.5.12 (Characterization of Smooth Sections)
Let π : E → M be a smooth vector bundle and U ⊆ M an open subset. Suppose
s1 , . . . , sr is a smooth frame for E over U . Then a section
eli
X
s= cj s j
j
We have already proven the case where s1 , . . . , sr is the frame of a trivialization of E over
©F
Proof
Suppose s = j cj sj is a smooth section of E over U . Fix a point p ∈ U and choose a
P
128
terms of the frame t1 , . . . , tr ,
X
s= bi ti
i
X
sj = aij ti ,
j
the coefficients
P b j, aj will all be smooth functions on V by the previous lemma. Next we
ou
i i
Zh
b1 c1
b = ... = A ... = Ac.
br cr
At each point of V , being the transition matix between two bases, the matrix A is invert-
ible. By Cramer’s rule, A−1 is a matrix of smooth functions on V . Hence c = A−1 b is a
column vector of smooth functions on V . This proves that c1 , . . . , cr are smooth functions
at p ∈ U . Since p is an arbitrary point of U , the coefficients cj are smooth functions on
U.
x
The converse holds similarly to the base case since the collection of smooth sections forms
a C ∞ (U )-module.
Remark 3.5.13 If one replaces “smooth” by “continuous” throughout, the discussion in
eli
this subsection remains valid in the continuous category.
to 1. We typically ask that the partition of unity is indexed by the the same set as an open
cover {Uα } and that the support of ρα is contained in Uα so that ρα vanishes outside of Uα .
The existence of smooth partitions of unity is an important technical tool in the theory
of smooth manifolds that distinguishes it from that of real-analytic or complex manifolds.
We construct smooth ump functions on any manifold and prove the existence of a smooth
partition of unity on a compact manifold. The general case is more technical and omitted.
If we have some objective locally defined for each Uα , we have a generic way of extending it
129
to all of M as a “weighted sum”. Conversely, we can decompose a global object on a manifold
into a locally finite sum of local objects.
Recall that R× denote the group fo nonzero real numbers under multiplication. Also, recall
ou
the support of a real-valued function f : M → R is defined to be
Zh
verification of smoothness.
The main challenge in constructing a smooth bump function is to obtain a smooth step
function. Consider the smooth function
(
exp(−1/t), t > 0
f (t) := .
0, t≤0
We seek a smooth step function g(t) by dividing f (t) by some positive function `(t). Such
a g(t) vanishes for t ≤ 0. The denominator should be a positive function that agrees with
f (t) for t ≥ 1 so that g(t) = 1 for t ≥ 1. We can take `(t) = f (t) + f (1 − t) and consider
x
f (t)
g(t) := .
f (t) + f (1 − t)
eli
`(t) > 0 for all t ∈ R by construction and g(t) is smooth since it is the quotient of two
smooth functions with a non-vanishing denominator.
Given 0 < a < b ∈ R, we make the linear change of variables to map [a2 , b2 ] → [0, 1]:
x − a2
x 7→ .
©F
b 2 − a2
Then g : R → [0, 1] given by
x − a2
h(x) := g
b 2 − a2
is a smooth step function that vanishes for x ≤ a2 and is 1 for x ≥ b2 . We then perform
another change of variables within h to make the function symmetric
k(x) := h(x2 ).
130
Finally, set
x 2 − a2
ρ(x) := 1 − k(x) = 1 − g .
b 2 − a2
This is a smooth bump function at 0 ∈ R that is identically 1 on [−a, a] and has support in
[−b, b]. For any q ∈ R, ρ(x − q) is a smooth bump functionat q.
ou
In order to extend this construction to Rn , consider
kxkr2 − a2
σ(x) := ρ(kxk) = 1 − g .
b 2 − a2
This is a smooth bump function at 0 ∈ Rn that is 1 on the closed ball B̄(0, a) and has
support in the closed ball B̄(0, b). Smoothness is a result of composing smooth functions.
Zh
Again, translating by q yields a smooth bump function at q ∈ Rn ,
σ(x − q).
Proposition 3.6.1
Let q ∈ M be a point in a manifold and U 3 q a neighborhood of q. There is a smooth
bump function at q supported in U .
Proof
x
Let V ⊆ U be a neighborhood of q and ϕ : V → Rn a coordinate map on V . Without
loss of generality, by translating if necessary, we can assume ϕ(q) = 0 ∈ Rn . Pick ε > 0
sufficiently small so that the open Euclidean ball B(0; 3ε) ⊆ ϕ(V ). Let σ : Rn → R be
the smooth bump function that is 1 on B̄(0, ε) and has support in B̄(0, 2ε). Note that
eli
supp σ ⊆ B(0; 3ε). Then the function σ ◦ ϕ : V → Rn → R is the desired bump function.
In general, a smooth function on an open subset U of a manifold M cannot be extended to a
smooth function on all M . An example is the secant function on the open interval (−π/2, π/2)
in R. However, this is possible if we relax the condition so that the global function agrees
with M only on some neighborhood of a point in U .
©F
Proof
Choose a smooth bump function ρ : M → R supported in U that is 1 on a neighborhood
131
V of p. Define (
ρ(q)f (q), q ∈ U
f˜(q) :=
0, q∈/U
ou
Since ρ ≡ 1 on V , the function f˜ agrees with f on V .
A collection {Ui } of subsets of a topological space S is said to be locally finite if every point
Zh
q ∈ S has a neighborhood that meets only finite many of the sets Ui . In particular, every
q ∈ S is contained in only finitely many of the Ui ’s.
Given an open cover {Uα }α∈A of M , we say that a partition of unity {ρα }α∈A is subordinate
x
to the open cover {Uα } if supp ρα ⊆ Uα for every α ∈ A.
The sum in (ii) makes sense as the locally finite condition ensures that it is a finite sum at
any given point.
eli
In general, consider {fα } a collection of smooth functions on a manifold M such that the
collection of its supports {supp fα } is locally finite. Then every q ∈ P
M has a neighborhood
Wq that intersects only finitely many supp fα . Thus on Wq , the sum α fα is actully a finite
sum. Thus shows that the functio is well-defined and smooth on the manifold M . We call
such a sum a locally finite sum.
©F
132
Lemma 3.6.3
If ρ1 , . . . , ρm are real-valued functions on a manifold M , then
!
X [
supp ρi ⊆ supp ρi .
i i
ou
Proof
Define
X
A := {p ∈ M : ρi (p) 6= 0}
i
Zh
X
supp ρi := Ā
i
[
⊆ Ai
i
[
= Ai
i
[
= supp ρi .
i
Here we used the fact that B ∪ C = B̄ ∪ C̄. A fact that can be proved using the definition
x
of the closure of B as the smallest closed set containing B.
Proposition 3.6.4
Let M be a compact manifold and {Uα }α∈A an open cover of M . There is a smooth
eli
partition of unity {ρα }α∈A subordinate to {Uα }α∈A .
Proof
For each q ∈ M , find an open set Uα 3 q and let Ψq be a smooth bump function at q
supported in Uα . Because Ψq (q) > 0, there is a neighborhood Wq of q on which Ψq > 0.
By the compactness of M , we can find a subcover
©F
Consider m
X
Ψ := Ψqi .
i=1
133
Define
Ψqi
ϕi :=
Ψ
for i ∈ [m]. We have i ϕi = 1 by construction. Also, this is a finite sum and hence
P
locally finite. Hence we already have a partition of unity and it remains to show that it
is subordinate to {Uα } by re-indexing.
Since Ψ > 0, ϕi (q) 6= 0 if and only if Ψqi (q) 6= 0 so that
ou
supp ϕi = supp Ψqi ⊆ Uα
for some α ∈ A. For each i ∈ [m], choose τ (i) ∈ A to be an index such that supp ϕi ⊆ Uτ (i) .
Group the functions {ϕi } by τ (i) and define
X
ρα := ϕi
Zh
i∈[m]:τ (i)=α
134
Vector fields are abundant in nature. For instance, the velocity vector field of a fluid flow,
the electric field of a charge, the gravitation field of a mass, and so on. The fluid flow model
is quite natural, as every smooth vector field may be viewed locally as the velocity vector
field of a fluid flow. The path traced out by a point under this flow is called the integral
curve of the vector field. Integral are curves whose velocity vector field is the restriction of
the given vector field to the curve. The theory of ODEs guarantee the existence of integral
curves.
ou
3.7.1 Smoothness of Vector Field
Zh
X ∂
Xp = ai (p) .
i
∂xi p
for v ∈ Tp M .
eli
Comparing coefficients of Xp
X ∂ X ∂
Xp = ai (p) = ci (Xp )
i
∂xi p i
∂xi p
on U .
135
Proof
We have proven a more general fact: For any smooth vector bundle π : E → M and P open
subset U ⊆ M . If s1 , . . . , sr is a smooth frame for E over U , then a section s = j cj sj
of E over U is smooth if and only if the cj ’s are smooth functions on U .
We let π be the tangent bundle and si = ∂/∂xi be the coordinate section to complete the
proof.
ou
We can now characterize the smoothness of a vector field on a manifold in terms of its
coefficients relative to coordinate frames.
Zh
ai of X = i ai ∂/∂xi relative to the frame ∂/∂xi are all smooth
P
Just as in the Euclidean case, a vector field X on a manifold M induces a linear map on the
algebra C ∞ (M ) of smooth functions on M . For f ∈ C ∞ (M ), define Xf to be the function
(Xf )(p) = Xp f.
We can now state an alternative characterization of a smooth vector field in terms of its
action as an operator on smooth functions.
x
Proposition 3.7.3 (Smoothness of a Vector Field in terms of Functions)
A vector field X on M is smooth if and only if for every smooth function f on M , the
function Xf is smooth on M .
eli
Proof
( =⇒ ) Suppose X is smooth so that on any chart (U, x1 , . . . , xn ) of M , the coefficients ai
of X = i ai ∂/∂xi are smooth. For any f ∈ C ∞ (M ), it follows that Xf = i ai ∂f /∂xi
P P
is smooth on U . Since M can be covered by charts, Xf is smooth on M .
©F
Each of the coordinate functions xk can be extended to a smooth function x̃k on M that
agrees with xk in a neighborhood V of p ∈ U . Thus on V ,
! !
X ∂ X ∂
X x̃k = ai i x̃k = ai i x k = ak .
i
∂x i
∂x
By assumption, each ak is smooth on V and in particular at p. But p ∈ M was arbitrary,
concluding the proof.
136
The proposition above shows that we can view a smooth vector field X as a linear operator
on C ∞ (M ). Similar to the Euclidean case, X is a derivation: for all f, g ∈ C ∞ (M ),
An alternative view of smooth vector fields on a manifold M can be taken as smooth sections
of the tangent bundle T M and as derivations on the algebra C ∞ (M ) of smooth functions.
ou
In fact, it can be shown that these two descriptions of smooth vector fields are equivalent.
Similary to smooth extensions of smooth functions, we can smoothly extend vector fields
using bump functions.
Proposition 3.7.4
Suppose X is a smooth vector field defined on a neighborhood U 3 p in a manifold
M . Then there is a smooth vector field X̃ on M that agrees with X on some (possibly
Zh
smaller) neighborhood of p.
Example 3.7.5
Let X be the vector field x2 d/dx on the real line R. We wish to determine the maximal
integral curve of X starting at x = 2.
©F
d d
x0 (t) = Xx(t) ⇐⇒ ẋ(t) = x2 .
dt dt
Thus x(t) satisfies the differential equation
dx
= x2
dt
137
subject to x(0) = 2. Solving the above by separation of variables yields
dx
= dt
x2
1
− =t+C
x
1
x=− .
ou
t+C
The initial conditions forces C = −1/2. Hence
2
x(t) = .
1 − 2t
The maximal interval containing 0 on which x(t) is defined is (−∞, 1/2).
Zh
This example shows that it may not be possible to extend the domain of an integral curve
to the entire real line.
Finding an integral curve of a vector field locally amounts to solving a system of first-
order ODE with initial conditions. Suppose we wish to find an integral curve c(t) of a
smooth vector field X on a manifold M in general. We first choose a coordinate chart
(U, φ) = (U, x1 , . . . , xn ) about p. In terms of the local coordinates,
x
X ∂
Xc(t) = ai (c(t))
i
∂xi c(t)
and
eli
X ∂
c0 (t) = ċi (t) ,
i
∂xi c(t)
where c (t) = x ◦ c(t) is the i-th component of c(t) in the chart (U, φ). The condition
i i
for each i ∈ [n]. This is an ODE with initial condition c(0) = p translating to
ci (0) = pi
Remark 3.7.6 We should think of elements of the tangent space as an infinitessimal direc-
tion and the differential of a map encoding how an infinitessimal direction in the domain
corresponds to an infinitessimal direction in the image of the map.
138
Also recall that partial derivatives simplify to the calculus derivative for maps between
Euclidean spaces.
By the existence and uniqueness of solutions to ODEs, the system above always has a unique
local solution.
Theorem 3.7.7
ou
Let V ⊆ Rn be open, p0 ∈ V , and f : V → Rn a smooth function. Then the
differential equation
dy
= f (y(t))
dt
with initial conditions y(0) = p0 has a unique smooth solution y : (a(p0 ), b(p0 )) → V
where (a(p0 ), b(p0 )) is the maximal open interval containing 0 on which y is defined.
Zh
Uniqueness above means that if z : (δ, ε) → V satisfies the same ODE, then (δ, ε) ⊆
(a(p0 ), b(p0 )) and z, y agree on their common domain.
This theorem guarantees the existence and uniquness of a maximal integral curve starting
at p for a vector field X on a chart U of a manifold and a point p ∈ U .
Next we study the dependence of an integral curve on its initial point. Again we begin with
the problem locally on Rn . The function y will now be a function of two arguments t, q, and
the condition for y to be an integral curve starting at q is
∂y
x
(t, q) = f (y(t, q))
∂t
with initial conditions y(0, q) = q.
The following theorem from ODE theory guarantees the smooth dependence of the solution
eli
on the initial point.
Theorem 3.7.8
Let V ⊆ Rn be open and f : V → Rn be smooth on V . For each p0 ∈ V , there is a
neighborhood W 3 p0 in V , some ε > 0, and a smooth function y : (−ε, ε) × W → V
such that
©F
∂y
(t, q) = f (y(t, q))
∂t
and y(0, q) = q for all (t, q) ∈ (−ε, ε) × W .
It follows from this theorem that if X is a any smooth vector field on a chart U with p ∈ U ,
then there is a neighorhood W 3 p in U , ε > 0, and a smooth map F : (−ε, ε) × W → U
such that for each q ∈ W , the function F (·, q) is an integral curve of X starting at q. In
particular, F (0, q) = q. We usually write Ft (q) = F (t, q).
139
Suppose s, t ∈ (−ε, ε) are such that both Ft (Fs (q)) and Ft+s (q) are defined. Then both
Ft (Fs (q)) and Ft+s (q) as functions of t are integral curves of X with initial point Fs (q). By
the uniqueness of the integral curve starting at a point,
The map F is called a local flow generated by X. For each q ∈ U , the function Ft (q) of t
is called a flow line of the local flow. Each flow line is an integral curve of X. If a local
ou
flow F is defined on R × M , it is called a global flow. Every smooth vector field has a local
flow about any point, but no necessarily a global flow. A vector field having a global flow is
called a complete vector field. If F is a global flow, then for every t ∈ R,
Ft ◦ F−t = F−t ◦ Ft = F0 = 1M
Zh
parameter group of diffeomorphisms of M .
The discussion above suggests the following formal definition.
F : (−ε, ε) × W → U,
Proposition 3.7.9
Let F : (−ε, ε) × W → U be a local flow about p ∈ U . The infinitessimal generator X
of F is a smooth vector field on W and each curve F (·, p) is an integral curve of X.
Proof
We wish to show that Xf is smooth for every smooth real-valued function f on an open
140
subset V ⊆ W . For any such f and p ∈ U , write c(t) = F (t, p) so that c0 (0) = Xp
(Xf )(p) = Xp f
= c0 (0)f
d
= c∗ f
dt 0
ou
d
= f (c(t))
dt 0
∂
= f (F (t, p)).
∂t (0,p)
Since compositions preserve smoothness, f (F (t, p)) is a smooth function of (t, p) and so
is its partial derivative with respect to t. Thus (Xf )(p) depends smoothly on p and X is
smooth.
Zh
We wish to show that XFt0 (p) = d/dtF (t0 , p) for all p ∈ W and t0 ∈ (−ε, ε). Define
q := Ft0 (p) and we show that d/dtF (t0 , p) = Xq . By the group law,
d
Xq f = f (F (t, q))
x
dt 0
d
= f (F (t0 + t, p))
dt 0
d
eli
= f (F (t, p))
dt t0
as desired.
©F
141
3.7.4 The Lie Bracket
ou
Suppose X, Y are smooth vector fields on an open subset U of a manifold M . We view X, Y
as derivation operators on C ∞ (U ). The map XY is an R-linear operator but does not satisfy
the Leibniz rule. If we consider XY − Y X however, this will be a derivation.
Zh
[X, Y ]p f := (Xp Y − Yp X)f.
i j
Then
X ∂
[X, Y ] = ck
k
∂xk
where
X ∂bk k
k i i ∂a
c = a i
−b i
.
i
∂x ∂x
142
Proof
Fix p ∈ U . Evaluating [X, Y ]p at the coordinate functions xk ’s yields
ck (p) = [X, Y ]p xk
= (Xp Y − Yp X)xk
= X p b k − Y p ak
ou
X ∂bk k
i i ∂a
= a i
−b i
.
i
∂x ∂x
Zh
(i) (bilinearity) [aX + bZ, Z] = a[X, Z] = b[Y, Z]
and [Z, aX + bY ] = a[Z, X] + b[Z, Y ]
(ii) (anticommutativity) [Y, X] = −[X, Y ]
(iii) (Jacobi identity) [X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y ]] = 0
In practice, we only concern ourselves with real Lie algebras, i.e. Lie algebras over R. From
hereonforth, a Lie algebra means a real Lie algebra.
Example 3.7.14
If M is a manifold, the vector space X(M ) of smooth vector fields on M is a real Lie
algebra with the Lie bracket as the bracket.
©F
Example 3.7.15
Let Kn×n be the vector space of all n × n matrices over a field K. Define for X, Y ∈ Kn×n ,
[X, Y ] = XY − Y X,
where XY is the matrix product. With this bracket, Kn×n becomes a Lie algebra.
More generally, if A is any algebra over a field K, then the product
[x, y] = xy − yx
143
makes A into a Lie algebra over K.
ou
Example 3.7.16
Let V be a Lie algebra over a field K. For each X ∈ V , define adX : V → V by
adX (Y ) := [X, Y ].
Zh
adX [Y, Z] = [X, [Y, Z]]
= [[X, Y ], Z] + [Y, [X, Z]]
= [adX Y, Z] + [Y, adX Z].
144
Definition 3.7.6 (Related Vector Field)
Let F : N → M be a smooth map between manifolds. A vector field X on N is
F -related to a vector field X̄ on M if for all p ∈ N ,
ou
Example 3.7.17
If F : N → M is a diffeomorphism and X is a vector field on N , then the pushforward
F∗ X on M is defined. By definition, X is F -related to the vector field F∗ X on M .
An equivalent condition of F -relatedness is as follows.
Proposition 3.7.18
Let F : N → M be a smooth map of manifolds. A vector feld X on N and a vector
Zh
field X̄ on M are F -related if and only if for all g ∈ C ∞ (M ),
X(g ◦ F ) = (X̄g) ◦ F.
Proof
X, X̄ are F -related if and only if
Proposition 3.7.19
eli
Let F : N → M be a smooth map of manifolds. If the smooth vector fields X, Y on N
are F -related to the smooth vector fields X̄, Ȳ , respectively, on M , then the Lie bracket
[X, Y ] on N is F -related to the Lie bracket [X̄, Ȳ ] on M .
Proof
For any g ∈ C ∞ (M ),
©F
145
©F
146
eli
x
Zh
ou
Chapter 4
ou
Lie Groups and Lie Algebras
Zh
A Lie group is a manifold equipped with smooth group operations. The invertible matrix
groups form important and interesting examples of Lie groups. The left translation by a
group element g is a diffeomorphism from the group to itself that maps the identity to g.
Thus the group locally looks the same around any point. To study the local structure of a
Lie group, it suffices to examine a neighborhood of the identity element. It is not surprising
that the tangent space at the identity of a Lie group should play a role.
The tangent space at the identity of a Lie group turns out to have a canonical bracket that
makes it into a Lie algebra. This encodes within it much information about the group.
The interplay of group theory, topology, and linear algebra makes the theory of Lie groups
x
and Lie algebras a particularly rich and vibrant branch of mathematics. Our humble goal
is to examine Lie groups as an important class of manifolds and Lie algebras as examples of
tangent spaces.
eli
147
Definition 4.1.1 (Lie Group)
A Lie group is a smooth manifold G that is also a group such that the two group
operations, multiplication and inverse are smooth.
µ:G×G→G µ(a, b) = ab
ι:G→G ι(a) = a−1 .
ou
for a ∈ G, we denote by `a (x) = ax the left multiplication operation by a, and by ra (x) = xa
the right multiplication operation by a. We also refer to left/right multiplication as left/right
translation.
Proposition 4.1.1
For an element a in a Lie group G, the left multiplication `a : G → G is a diffeomorphism.
Zh
Proof
Consider the inclusion map I : G → G × G given by I(x) = (a, x). This is certainly
smooth. Then the composition
`a = I ◦ µ
is smooth.
F ◦ `h = `F (h) ◦ F
for all h ∈ H. Recall that group homomorphisms always map the identity to itself.
©F
We use capital letters to denote matrices, but generally lowercase letters to denote their
entries.
is a Lie group.
148
Example 4.1.3 (Special Linear Group)
The special linear group SL(n, R) is the subgroup of GL(n, R) consisting of matrices with
determinant 1. We know that SL(n, R) is a regular submanifold of dimension n2 − 1.
Recall that smooth maps f : N → M whose image lie in a regular submanifold S ⊆
induces a smooth map f˜ : N → S. Thus multiplication and inverse operations from
GL(n, R) induce smooth multiplication and inverse operations on SL(n, R).
ou
An analogous argument proves that the complex special linear group SL(n, C) is also a
Lie group.
Zh
set theorem, O(n) is a regular submanifold of GL(n, R). By a similar argument to the
previous example, it is hence a Lie group.
The constant rank level set theorem is more general than the regular level set theorem, at
the cost that we do not directly know the co-dimension of the regular submanifold. We wish
to directly apply the regular level set theorem to determine the dimension of O(n).
While it is true that f also maps GL(n, R) → GL(n, R) or Rn×n , we cannot hope for the
differential to be surjective. This illustrates a general principle: for the differential to be
©F
For any matrix X ∈ Rn×n , we know that there is a curve c(t) in GL(n, R) with c(0) = A
149
and c0 (0) = X. Then
d
f∗,A (X) = f (c(t))
dt t=0
d
= c(t)T c(t)
dt t=0
ou
= (c0 (t)c(t) + c(t)T c0 (t)) matrix product rule
t=0
= X T A + AT X.
Fix A ∈ O(n) and B ∈ Sn . To show surjectivity, it suffice to find some X ∈ Rn×n such that
X T A + AT X = B.
Zh
This equation has the solution
1
X = (AT )−1 B.
2
Thus f∗,A : TA GL(n, R) → Sn is surjective for all A ∈ O(n) and O(n) is a regular level set of
f . By the regular level set theorem, O(n) is a regular submanifold of GL(n, R) of dimension
n2 − n
n2 − dim Sn = .
2
The group operations on H must be the restrictions of the operations inherited from G.
However, since a Lie group is an immersed submanifold instead of a regular submanifold, it
need not have the relative topology. Let ι : H → G be the inclusion map. The composition
of smooth maps
µ ◦ (ι × ι) : H × H → G × G → G
is smooth. If we took H to be a regular submanifold, then the multiplication map H×H → H
and inverse map H → H would automatically be smooth and condition (ii) is redundant.
150
Proposition 4.1.6
If H is a subgroup and regular submanifold of a Lie group G, then it is a Lie subgroup
of G.
A subgroup such as in the proposition above is called an embedded Lie subgroup, as the
inclusion map ι : H → G of a regular submanifold is an embedding.
ou
Example 4.1.7
SL(n, R) and O(n) are embedded Lie subgroups of GL(n, R).
We now state a powerful theorem without proof.
Zh
Example 4.1.9
SL(n, R) and O(n) are the zero sets of polynomial equations on GL(n, R) and hence are
closed subsets of GL(n, R). It follows that they are embedded Lie subgroups of GL(n, R).
In order to compute the differential of a map on a subgroup of GL(n, R), we need a curve of
nonsingular matrices. Since the matrix exponential is always nonsingular, it is suitable for
this purpose.
x
The vector space Rn×n ' Rn of n × n matrices can be given the Euclidean norm
2
X 2
x2ij .
eli
kXk =
ij
For this formula to make sense, we need to check that it converges in the normed linear space
Rn×n .
kvwk ≤ kvkkwk.
151
Matrix multiplication makes Rn×n into a normed algebra.
Proposition 4.1.10
For X, Y ∈ Rn×n ,
kXY k ≤ kXkkY k.
Proof
ou
Cauchy-Schwartz inequality.
In a normed algebra, multiplication distributes over a finite sum. The distributivity of
multiplication over an infinite sum requires proof.
Proposition 4.1.11
Let V be a normed algebra.
(i) If a ∈ and (sm ) is a sequence in V that converges to s, then asm converges to as.
Zh
(ii) If a ∈ V and k≥0 bk is a convergent series in V , then a k bk = k abk .
P P P
Recall a series k ak in a normed vector space V is said to converge absolutely if the series
P
true that
eA+B = eA eB .
Proposition 4.1.12
If A, B ∈ Rn×n commute, then
eA+B = eA eB .
152
Proposition 4.1.13
For X ∈ Rn×n ,
d tX
e = XetX = etX X.
dt
Proof
Since each (i, j)-th entry of the series for the exponential function et X is a convergent
ou
power series in t, it is justified to differentiate term by term.
d tX d 1 2 2 1 3 3
e = Id +tX + t X + t X
dt dt 2! 3!
1
= X + tX 2 + t2 X 3
2!
1 2 2
Zh
= X Id +tX + t X + . . .
2!
t
= Xe X.
Lemma 4.1.14
(i) For any X, Y ∈ Rn×n , tr(XY ) = tr(Y X).
(ii) For X ∈ Rn×n and A ∈ GL(n, R), tr(AXA−1 ) = tr(X).
©F
Proposition 4.1.15
The trace of a real or complex matrix is equal to the sum of its (complex) eigenvalues.
Proposition 4.1.16
For any X ∈ Rn×n , det eX = etr X .
153
Proof
Suppose first that X is upper triangular. Then the diagonal entries of eX is exii for i ∈ [n].
Thus Y
det eX = exii = etr X .
i
If X is not upper triangular, we can find a nonsingular matrix such that AXA−1 is
ou
triangular. Thus
X 1 X 1
AXA−1 −1 k
e = (AXA ) = A X A−1 = AeX A−1 .
k
k≥0
k! k≥0
k!
But then
AXA−1 AXA−1
X X −1
= etr = etr X
det e = det Ae A = det e
Zh
by the special case.
It follows that the matrix exponential eX is never singular regardless of the X, as its deter-
minant is strictly positive.
This enables us to write down an explicit curve in GL(n, R) with given initial point and
given iniital velocity. For example, c(t) = etX : R → GL(n, R) is a curve in GL(n, R) with
initial point Id and initial velocity X, since
c(0) = e0X = I
and
x
d
c0 (0) = etX = XetX = X.
dt t=0 t=0
Similarly, c(t) = AetX is a curve in GL(n, R) with initial point A and initial velocity AX.
eli
The tangent space TId GL(n, R) at the identity matrix Id is the vector space Rn×n and the
tangent space T1 R is R. Thus we can view the differential of the determinant map
©F
det : Rn×n → R
∗,Id
Proposition 4.1.17
For any X ∈ Rn×n ,
det(X) = tr X.
∗,Id
154
Proof
Choose the matrix exponential c(t) = etX so that c(0) = Id and c0 (0) = X. Then
d
det etX
det(X) =
∗,Id dt t=0
d
= et tr X
ou
dt t=0
= (tr X)et tr X
t=0
= tr X.
Zh
In a Lie group G, left translation by an element e ∈ G is a diffeomorphism that maps a
neighborhood of the identity to a neighborhood of g. Thus all the local information about
the group is concentrated in a neighborhood of the identity, and the tangent space at the
identity is especially important.
We can give the tangent space Te G a Lie bracket [, ] so that it becomes a Lie algebra, called
the Lie algebra of the Lie group. Our goal is to define the Lie algebra and identity the Lie
algebras of a few important groups.
The Lie bracket on the tangent space Te G is defined using a canonical isomorphism between
x
the tangent space at the identity and the vector space of left-invariant vector fields on G.
With respect to this Lie bracket, the differential of a Lie group homomorphism becomes a
Lie algebra homomorphism. We thus obtain a functor from the category of Lie groups and
eli
Lie group homomorphisms to the category of Lie algebras and Lie algebra homomorphisms.
The existence of a smooth multiplication and smooth inverse makes a Lie group a very special
©F
Thus if we describe the tangent space Te G, then `g∗ Te G will give a description of the tangent
space Tg G at any g ∈ G.
155
Example 4.2.1 (Tangent space to GL(n, R) at Id)
We know Tg GL(n, R) ' Rn×n . We also identified the isomorphism `g∗ : TId GL(n, R) →
Tg GL(n, R) as left multiplication by g : X 7→ gX.
ou
SL(n, R), this curve has constant determinant 1. We now take the derivative at t = 0.
d
0= det(c(t))
dt
t=0
d
= (det ◦c)∗
dt 0
Zh
d
= det c∗
∗,I dt 0
0
= det(c (0))
∗,I
= det(X)
∗,I
Thus the tangent space is contained in the subspace of trace 0 matrices. But this subspace
has dimension n2 − 1 = dim TId SL(n, R) and the two spaces must be equal.
x
Proposition 4.2.3
TId SL(n, R) can be identified with the subspace of trace 0 n × n matrices.
eli
Example 4.2.4 (Tangent Space to O(n) at Id)
Let X ∈ TId O(n) and choose a curve in O(n) defined on a small interval about 0 such
that c(0) = I and c0 (0) = X. Since c(t) ∈ O(n),
c(t)T c(t) = Id .
Differentiating both sides with respect to t using the matrix product rule yields
©F
156
and these vector spaces must be equal.
Proposition 4.2.5
The tangent space TId O(n) can be identified with the n × n skew-symmetric matrices.
ou
Let X be any not necessarily smooth vector field on a Lie group G. Fro any g ∈ G, since
left multiplication `g : G → G is a diffeomorphism, the pushforward `g∗ X is a well-defined
vector field on G. We say that the vector field X is left-invariant if
`g∗ (X) = X
Zh
for every g ∈ G. This means that for any h ∈ G,
In other words, X is left-invariant if and only if it is `g -related to itself for all g ∈ G. Thus
a left-invariant vector field X is completely determined by its value Xe at the identity, since
Xg = `g∗ (Xe ).
We say that à is the left-invariant vector field on G generated y A ∈ Te G. Let L(G) be the
vector space of all left-invariant vector fields on G. Then there is a bijective correspondence
Te G ↔ L(G)
Xe ←[ X
A 7→ Ã.
157
Example 4.2.6 (Left-Invariant Vector Fields on R)
On the Lie group R, the group operation is addition and the identity element is 0. Thus
“left multiplication” `g is actually left addition
`g (x) = g + x.
Let us compute `g∗ (d/dx|0 ). Since `g∗ ∗ (d/dx|0 ) is a tangent vector at g, it is a scalar
multiple of d/dx|g :
ou
d d
`g∗ =a .
dx 0 dx g
In order to compute a, we can evaluate both sides at the function f (x) = x to see that
a = 1. Thus
d d
`g∗ = .
dx 0 dx g
Zh
This shows that d/dx is a left-invariant vector field on R. Therefore, the left-invariant
vector fields on R are constant multiples of d/dx.
X ∂
aij ↔ [aij ].
x
ij
∂xij g
B̃g = (`g )∗ B ↔ gB
Proposition 4.2.8
©F
Proof
We show that for any f ∈ C ∞ (G), the function Xf is also smooth. Choose a smooth
curve c : I → G on some interval about 0 such that c(0) = e and c0 (0) = Xe . If g ∈ G,
then gc(t) is a curve starting at g with initial vector Xg , since gc(0) = ge = g and
(gc)0 (0) = `g∗ c0 (0) = `g∗ Xe = Xg .
158
Then
d
(Xf )(g) = Xg f = f (gc(t)).
dt t=0
The function F (g, t) := f (gc(t)) is a composition of smooth functions and is thus smooth:
Id ×c µ f
G×I −−−→ G×G −→ G −→ R
7→ 7→ 7→
ou
(g, t) (g, c(t)) gc(t) f (gc(t)).
Its partial derivative ∂F (g, t)/∂t with respect to t is therefore also smooth. But then
∂F (g, t)/∂t|t=0 = (Xf )(g) is thus also smooth. This shows that X is indeed a smooth
vector field on G.
This proposition shows that the vector space L(G) of left-invariant vector fields on G is a
subspace of the vector space X(G) of all smooth vector fields on G.
Zh
Proposition 4.2.9
If X, Y are left-invariant vector fields on G, then so is [X, Y ].
Proof
For any g ∈ G, X is `g -related to itself, and Y is `g -related to itself. But then we know
that [X, Y ] is `g -related to itself.
has a natural notion of pushforward, given by the differential of a Lie group homomorphism.
The linear isomorphism ϕ : Te G ' L(G) allows us to define a Lie bracket on Te G and to
push forward left-invariant vector fields under a Lie group homomorphism.
We begin with the Lie bracket on Te G. Given A, B ∈ Te G, we first map them via ϕ to the
left-invariant vector fields Ã, B̃, take the Lie bracket [Ã, B̃] = ÃB̃ − B̃ Ã, and then map it
back to Te G via ϕ−1 . Thus the definition of the Lie algebra [A, B] ∈ Te G should be
159
Proposition 4.2.10
If A, B ∈ Te G and Ã, B̃ are the left-invariant vector fields they generate, then
^
[Ã, B̃] = [A, B].
Proof
ou
Applying ()
e to both sides of the equation [A, B] = [Ã, B̃]e yields
^
[A, ^
B] = [Ã, B̃]e = [Ã, B̃],
since ()
e and ()e are inverse to each other.
With the Lie bracket [, ], the tangent space Te G becomes a Lie algebra, called the Lie algebra
of the Lie group G. As a Lie algebra, Te G is usually denoted by g.
4.2.4
Zh
The Lie Bracket on gl(n, R)
For GL(n, R), the tangent space at Id can be identified with the vector space of n × n real
matrices. We make the identification
X
ij
aij
∂
∂xij Id
↔ [aij ].
The tangent space TId GL(n, R) with its Lie algebra structure is denoted by gl(n, R). Let Ã
x
be the left-invariant vector field on GL(n, R) generated by some A ∈ gl(n, R). Then on the
Lie algebra gl(n, R) we have the Lie bracket [A, B] = [Ã, B̃]Id coming from the Lie bracket
of left-invariant vector fields.
eli
In the following proposition, we identify the Lie bracket in terms of matrices.
Proposition 4.2.11
Let A, B ∈ TId GL(n, R). If
X ∂
©F
then X
cij = aik bkj − bik akj .
k
[A, B] = AB − BA.
160
Proof
We evaluate both sides of
X ∂
[A, B] = cij
ij
∂xij Id
ou
cij = [Ã, B̃]Id xij
= ÃId B̃xij − B̃Id Ãxij
= AB̃xij − B Ãxij .
In order to compute B̃xij , recall that the left-invariant vector field B̃ on GL(n, R) is given
by
X ∂
Zh
B̃g = (gB)ij .
ij
∂x ij g
Hence
B̃g xij = (gB)ij
X
= gik bkj
k
X
= bkj xik (g).
k
Since this formula holds for all g ∈ GL(n, R), the function B̃xij must be
x
X
B̃xij = bkj xik .
k
But then
eli
!
X ∂ X
AB̃ij = apq bkj xik
pq
∂xpq Id k
X
= apq bkj δip δkq
pqk
X
= aik bkj
©F
k
= (AB)ij
and
B Ãxij = (BA)ij .
It follows that
cij = (AB − BA)ij
as desired.
161
4.2.5 The Pushforward of Left-Invariant Vector Fields
ou
Let F : H → G be a Lie group homomorphism. A left-invariant vector field X on H is
generated by its value A = Xe ∈ Te Hat the identity, so that X = Ã. Since the Lie group
homomorphism F maps the identity of H to the identity of G, its differential F∗,e at the
identity is a linear map from Te H → Te G. The commutative diagram below shows the
existence of an induced linear map F∗ : L(G) → L(G) on left-invariant vector fields as well
as a way to define it.
Zh
F∗,e
Te H Te G A F∗,e A
' '
L(H) L(G) Ã ^
(F∗,e A)
Definition 4.2.1
x
Let F : H → G be a Lie group homomorphism. Define F∗ : L(G) → L(H) by
^
F∗ (Ã) := (F∗,e A)
eli
for all A ∈ Te H.
Proposition 4.2.12
If F : H → G is a Lie group homomorphism and X is a left-invariant vector field on H,
then X is F -related to the left-invariant vector field F∗ X on G.
©F
Proof
Fix h ∈ H. It suffices to check that
162
while the RHS is equal to
But F is a Lie group homomorphism, thus we have F ◦ `h = `F (h) ◦ F and the two sides
ou
are equal.
Thus if F : H → G is a Lie group homomorphism and X is a left-invariant vector field on
H, we will call F∗ X the pushforward of X under F .
Zh
Proposition 4.2.13
If F : H → G is a Lie group homomorphism, then its differential at the identity
F∗ := F∗,e : Te H → Te G
Proof
x
By a previous proposition, the vector field F∗ à on G is F -related to the vector field à on
H, and the vector field F∗ B̃ is F -related to B̃ on H. Hence the bracket [F∗ Ã, F∗ B̃] on G
is F -related to the bracket [Ã, B̃] on H. In other words,
eli
F∗ ([Ã, B̃]e ) = [F∗ Ã, F∗ B̃]F (e)=e .
^ ^
[F∗ Ã, F∗ B̃]e = [(F∗ A), (F ∗ B)]e
©F
= [F∗ , F∗ B].
163
proposition,
ι∗ ([X, Y ]Te H ) = [ι∗ X, ι∗ Y ]Te G .
Thus if Te H is identified with a subspace of Te G via ι∗ , then the bracket on Te H is simply
the restriction of the bracket on Te G to Te H. Thus the Lie algebra of a Lie subgroup H may
be identified with a Lie subalgebra of the Lie algebra of G.
We typically denote the Lie algebras of the classical groups by gothic letters. The Lie algebras
ou
of GL(n, R), SL(n, R), O(n), U (n) are denoted by gl(n, R), sl(n, R), o(n), u(n), respectively.
Moreover, the Lie algebra structurse on sl(n, R), o(n), u(n) are given by
[A, B] = AB − BA
as on gl(n, R).
Remark 4.2.14 A fundamental theorem in Lie theory asserts the existence of a bijective
Zh
correspondence between the connected Lie subgroups of a Lie group G and the Lie subal-
gebras of its Lie algebra g. It is because of our desire for such a correspondence that a Lie
subgroup of a Lie group is defined to be a subgroup that is also an immersed submanifold
rather than a regular submanifold.
x
eli
©F
164
Chapter 5
ou
Differential Forms
Zh
5.1 Differential 1-Forms
(df )p (Xp ) = Xp f.
165
We may also write df |p for the value of the 1-form df at p. This is in parallel to the notation
for tangent vector d/dt|p .
Recall our other notion of the differential f∗ for a smooth function f : N → M as a linear
function between tangent spaces.
Proposition 5.1.1
If f : M → R is a smooth function, then for p ∈ M and Xp ∈ Tp M ,
ou
d
f∗ (Xp ) = (df )p (Xp ) .
dt f (p)
Proof
Evaluate both sides of
Zh
d
f∗ (Xp ) = a
dt f (p)
at x.
This proposition shows that under the canonical identification
d
a ↔ a,
dt f (p)
f∗ is the same as df . Hence we are justified in calling both of them the differential of f .
In terms of df , a smooth function f : M → R has a critical point at p ∈ M if and only if
x
(df )p = 0.
Proposition 5.1.2
At each p ∈ U , the covectors (dx1 )p , . . . , (dxn )p form a basis for the cotangent space
©F
Proof
The proof is identical to the Euclidean case:
!
∂
(dxi )p = δji .
∂xj p
166
We can thus write every 1-form ω on U as a linear combination
X
ω= ai dxi ,
i
ou
X
df = ai dxi .
i
We can isolate ai by the usual trick of evaluating at both sides on ∂/∂xj to see that
∂f
aj = .
∂xj
Zh
Thus we have the following local expression for df :
X ∂f
df = i
dxi .
i
∂x
The underlying set of the cotangent bundle T ∗ M of a manifold M is the (disjoint) union of
the cotangent spaces at all points of M :
x
G
T ∗ M := Tp∗ M.
p∈M
Just as in the case of the tangent bundle, there is a natural map π : T ∗ M → M given by
eli
π(α) = p if α ∈ Tp∗ M . Mimicking the construction of the tangent bundle, we give T ∗ M
a topology as follows: If (U, φ) = (U, x1 , . . . , xn ) is a chart on M and p ∈ U , then each
α ∈ Tp∗ M can be written uniquely as a linear combination
X
α= ci (α)dxi |p .
i
©F
φ̃ : T ∗ U → φ(U ) × Rn
α 7→ (φ(p), c1 (α), . . . , cn (α)) = (φ ◦ π, c1 , . . . , cn )(α).
167
collection of subsets of T ∗ M to be a basis. We give T ∗ M the topology generated by the basis
B. As for the tangent bundle, with the maps φ̃ = (x1 ◦ π, . . . , xn ◦ π, c1 , . . . , cn ) as coordinate
maps, T ∗ M becomes a smooth manifold, and the projection map π : T ∗ M → M becomes
a vector bundle of rank n over M , justifying the “bundle” in the name “cotangent bundle”.
If x1 , . . . , xn are coordinates on U ⊆ M , then π ∗ x1 , . . . , π ∗ xn , c1 , . . . , cn are coordinates on
π −1 U ⊆ T ∗ M . Properly speaking, the cotangent bundle of a manifold M is the triple
(T ∗ M, M, π), while T ∗ M and M are the total space and base space of the cotangent bundles
ou
respectively. By abuse of language, it is customary to call T ∗ M the cotangent bundle of M .
In terms of the cotangent bundle, a 1-form on M is simply a section of the cotangent bundle
T ∗ M , ie it is a map ω : M → T ∗ M such that π ◦ ω = IdM . We say that a 1-form ω is smooth
if it is smooth as a map M → T ∗ M .
Zh
a manifold of dimension 2n. Remarkably, on T ∗ M there is a 1-form λ, called the Liouville
form (Poincaré form), defined independently of charts as follows.
A point in T ∗ M is a covector ωp ∈ Tp∗ M at some point p ∈ M . If Xωp is a tangent vector
to T ∗ M at ωp , then the pushforward π∗ (Xωp ) is a tangent vector to M at p. Thus one
can pair up ωp and π∗ (Xωp ) to obtain a real number ωp (π∗ (Xωp )). Define
The cotangent bundle and the Liouville form on it play an important role in classical
mechanics.
x
5.1.4 Characterization of Smooth 1-Forms
eli
We define a 1-form ω on a manifold M to be smooth if ω : M → T ∗ M is smooth as a
section of the cotangent bundle π : T ∗ M → M . The set of all smooth 1-forms on M has the
structure of a vector space, denoted by Ω1 (M ). In a coordinate chart (U, φ) = (U, x1 , . . . , xn )
on M , the value of the 1-form ω at p ∈ U is a linear combination
X
ωp = ai (p)dxi |p .
©F
168
Comparing the coefficients in
X X
ωp = ai (p)dxi |p = ci (ωp )dxi |p ,
i i
ou
is also true.
Lemma 5.1.4
Let (U, φ) = (U, x1 , . . . , xn ) be a chart on a manifold M . A 1-form ω = i ai dxi on
P
U is smooth if and only if the coefficient functions ai are all smooth.
Proof
Zh
Recall the characterization of smooth sections over U which states that a section over U
is smooth if and only if its coefficient functions with respect to any smooth frame over U
is smooth. This is simply a special case with the cotangent bundle as the vector bundle
and the coordinate 1-forms dxj as the smooth frame.
Proof
On any chart (U, x1 , . . . , xn ) on M ,
©F
X ∂f
df = i
dxi .
i
∂x
169
Proposition 5.1.6 (Linearity of 1-Forms over Functions)
Let ω be a 1-form on a manifold M . If f is a function and X a vector field on M , then
ω(f X) = f ω(X).
Proof
At each p ∈ M ,
ou
ω(f X)p := ωp (f (p)Xp ) = f (p)ωp (Xp ) =: (f ω(X))p .
Zh
Proof
( =⇒ ) Let ω be a smooth 1-form and X a smooth vector field on M . On any chart
(U, x1 , . . . , xn ) on M , we have ω = i ai dxi and X = j bj ∂x∂ j . But then by the linearity
P P
of 1-forms over functions,
! !
X X ∂ X X
ω(X) = ai dxi bj j = ai bj δji = ai b i .
i j
∂x i,j i
about p so that ω = i ai dx on U . For any j ∈ [n], we can extend the smooth vector
field X := ∂/∂xj on U to a smooth vector field X̄ on M that agrees with ∂/∂xj in a
eli
neighborhood Vpj of p in U . Restricted to Vpj ,
!
X
i ∂
ω(X̄) = ai dx = aj .
i
∂xj
170
5.1.5 Pullback of 1-Forms
ou
Definition 5.1.2 (Codifferential)
The codifferential (dual of the differential),
pulls back a covector at F (p) from M to N . This means that if ωF (p) ∈ TF∗ (p) M is a
Zh
covector at F (p) and Xp ∈ Tp N is a tangent vector at p, then
We call F ∗ (ωF (p) ) the pullback of the covector ωF (p) by F . Thus the pullback of covectors is
simply the codifferential.
Unlike vectgor fields which in general cannot be pushed forward under a smooth map, every
x
covector field can be pulled back by a smooth map. If ω is a 1-form on M , its pullback F ∗ ω
is the 1-form on N defined pointwise as expected:
(F ∗ ω)p := F ∗ (ωF (p) )
eli
for p ∈ N . Thus for Xp ∈ Tp N ,
(F ∗ ω)p (Xp ) = ωF (p) (F∗ (Xp )).
Having defined the pullback of a 1-form, we turn to the natural question of whether this
operation preserves smoothness. First, we establish three commutation properties of the
©F
171
Proof
Fix p ∈ N and Xp ∈ Tp N . We check that
ou
(F ∗ dh)p (Xp ) = (df )F (p) (F∗ Xp )
= (F∗ Xp )h
= Xp (h ◦ F ).
Zh
= Xp (h ◦ F ).
We proceed to check that pullbacks of functions and 1-forms respect addition and scalar
multiplication.
(ii)
©F
172
Proposition 5.1.10 (Pullback of a Smooth 1-Form)
The pullback F ∗ ω of a smooth 1-form ω on M under a smooth map F : N → M is a
smooth 1-form on N .
Proof
Fix p ∈ N . It suffices to check that F ∗ ω is smooth at p. Choose a chart (V, y 1 , . . . , y n )
on M about F (p). By the continuity (smoothness) P of Fi , there is a chart∞ (U, x , . . . , x )
ou
1 n
Zh
X
= (ai ◦ F )d(F ∗ y i )
i
X
= (ai ◦ F )d(y i ◦ F )
i
X X ∂F i
= (ai ◦ F ) j
dxj
i j
∂x
X ∂F i
= (ai ◦ F ) j dxj .
i,j
∂x
173
for every p ∈ S and v ∈ Tp S.
Proposition 5.1.12
If ι : S → M is the inclusion map of an immersed submanifold S and ω is a 1-form on
M , then
ι∗ ω = ω|S .
ou
Proof
For p ∈ S and v ∈ Tp S,
Zh
For simplicity of notation, we sometimes write ω to mean ω|S .
Thus
∂ ∂
X = −y +x
∂x ∂y
is a smooth vector field on the unit circle S 1 .
x
This notation means that if x, y are the standard coordinates on R2 and ι : S 1 → R2 is
the inclusion map, then at a point p = (x, y) ∈ S 1 , one has
eli
∂ ∂
ι∗ Xp = −y +x ,
∂x p ∂y p
Example 5.1.14
©F
ω(X) ≡ 1
on the unit circle S 1 where X is the velocity vector field of the unit circle.
174
This is in contrast to the situation for vector fields where ι∗ (∂/∂ x̄p ) 6= ∂/∂x|p .
ou
h∗ (−ydx + xdy) = −(h∗ y)d(h∗ x) + (h∗ x)d(h∗ y)
= −(sin t)d(cos t) + (cos t)d(sin t)
= sin2 tdt + cos2 dt
= dt.
Zh
Similar to the Euclidean setting, we now generalize the construction of 1-forms on a manifold
to that of k-forms. In parallel to the construction
Vkof the tangent and cotangent bundles on a
manifold, we construct the k-th exterior power (T M ) of the tangent bundle. This yields
∗
aVnatural notion of smoothness of differential forms as smooth sections of the vector bundle
k
(T ∗ M ). The pullback and wedge product of differential forms are defined pointwise. We
consider left-invariant forms on a Lie group as examples of differential forms.
f : V × · · · × V → R.
eli
We say that the k-tensor f is alternating if for any permutation σ ∈ Sk ,
Note that all 1-tensors are alternating. An alternating k-tensor on V is also called a k-
covector on V .
©F
For any vector space V , denote Vk by ∨Ak (V ) the vector space of alternating k-tensorsVon V.
Another common notation is (V ). There is a purely algebraic construction of k
(V ),
called the k-th exterior power of the vector space V , with the property that k (V ∨ ) ' Ak (V ).
V
We ignore this construction for simplicity.
We apply the function Ak () to the tangent space Tp M of a manifold M at a point p. The
vector space Ak (Tp M ), typically denoted k (Tp∗ M ), is the space of all alternating k-tensors
V
on the tangent space Tp M . A k-covector field on M is a function ω that assigns to each
175
p ∈ M a k-covector ωp ∈ k (Tp∗ M ). A k-covector field is also called a differential k-form, a
V
differential form of degree k, or simply a k-form. A top form on a manifold is a differential
form whose degree is the dimension of the manifold.
If ω is a k-form on a manifold M and X1 , . . . , Xk are vector fields on M , then ω(X1 , . . . , Xk )
is the function on M defined by
ou
Proposition 5.2.1 (Multilinearity of a Form over Functions)
Let ω be a k-form on a manifold M . For any vector fields X1 , . . . , Xk and any function
h on M ,
ω(X1 , . . . , hXi , . . . , Xk ) = hω(X1 , . . . , Xi , . . . , Xk ).
Zh
Example 5.2.2
Let (U, x1 , . . . , xn ) be a coordinate chart on a manifold. At each p ∈ U , a basis for the
tangent space Tp U is ∂/∂x1 |p , . . . , ∂/∂xn |p . Recall the dual basis for the cotangent space
Tp∗ U is
(dx1 )p , . . . , (dxn )p .
As p varies over U , we get differential 1-forms dx1 , . . . , dxn on U .
Recall from the general theory of alternating k-tensors that a basis for (Tp∗ U ) is
Vk
x
(dxi1 )p ∧ · · · ∧ (dxik )p ,
In this expression, the coefficients ai1 ...ik are functions on U as they vary with the point p.
©F
to indicate the set of all strictly ascending multi-indices between 1 and n of length k, and
write X
ω= aI dxI ,
I∈Jk,n
176
5.2.2 Local Expression for a k-Form
ou
(
1, I = J,
dxI (∂j1 , . . . , ∂jk ) = δJI =
0, I 6= J.
Zh
Proposition 5.2.3 (A Wedge of Differentials in Local Coordinates)
Let (U, x1 , . . . , xn ) be a chart on a manifold and f 1 , . . . , f k smooth functions on U . Then
X ∂(f 1 , . . . , f k )
df 1 ∧ · · · ∧ df k = i1 , . . . , x ik )
dxi1 ∧ · · · ∧ dxik .
I∈J
∂(x
k,n
Proof
On U ,
X
df 1 ∧ · · · ∧ df k = cJ dxj1 ∧ · · · ∧ dxjk
x
J∈Jk,n
for some functions cJ . For the LHS, recall that the wedge product of functions applied k
vectors satisfies
eli
i
1 k ∂f
(df ∧ · · · ∧ df )(∂i1 , . . . , ∂ik ) = det
∂xij
∂(f 1 , . . . , f k )
= .
∂(xi1 , . . . , xik )
X X
cJ dxJ (∂i1 , . . . , ∂ik ) = cJ δIJ = cI .
J J
If (U, x1 , . . . , xn ) and (V, y 1 , . . . , y n ) are two overlapping charts on a manifold, then on the
intersection U ∩ V , the proposition above yields the transition formula for k-forms:
X ∂(y j1 , . . . , y jk )
dy J = i 1 , . . . , x ik )
dxI .
I∈J
∂(x
k,n
177
Two cases of the proposition above are of special interest:
Corollary 5.2.3.1
Let (U, x1 , . . . , xn ) be a chart on a manifold and f, f 1 , . . . , f n ∈ C ∞ (U ). Then
(i) (1-forms) df = i (∂f /∂xi )dxi
P
ou
Proposition 5.2.4 (Transition Formula for a 2-Form)
If (U, x1 , . . . , xn ) and (V, y 1 , . . . , y n ) are two overlapping coordinate charts on M , then
a smooth 2-form ω on U ∩ V has two local expressions
X X
ω= aij dxi ∧ dxj = bk` dy k ∧ dy ` .
i<j k<`
Zh
Then
X ∂(y k , y ` )
aij = bk` .
k<`
∂(xi , xj )
Proof
By an earlier remark,
X ∂(y k , y ` )
dy k ∧ dy ` = dxi ∧ dxj .
i<j
∂(xi , xj )
So that
x
X X X ∂(y k , y ` )
ω= aij dxi ∧ dxj = bk` dxi ∧ dxj .
i<j k<` i<j
∂(xi , xj )
Let M be a manifold of dimension n. We mimic the construction of the tangent and cotangent
bundles and form the set
©F
k
^ k
G ^ G
(T ∗ M ) := (Tp∗ M ) = Ak (Tp M )
p∈M p∈M
of all alternating k-tensors at all points of the manifold M . This set is called V
the k-th exterior
power of the cotangent bundle. There is a canonical projection map π : k
(T ∗ M ) → M
given by π(α) = p for α ∈
Vk ∗
(Tp M ).
178
If (U, φ) is a coordinate chart on M , then there is a bijection
k k
(Tp∗ U ) ' φ(U ) × R(k )
^ [^ n
∗
(T U ) =
p∈U
k
^
α∈ (Tp∗ U ) 7→ (φ(p), {cI (α)}I ),
ou
where α = I cI (α)dxI |p ∈ k (Tp∗ U ) and I = (1 ≤ i1 < · · · < ik ≤ n). Hence we can given
P V
Zh
If E → M is a smooth vector bundle, then the vector space of smooth sections of E is denoted
by Γ(E) or Γ(M, E). The vector space of all smooth k-forms on M is usually denoted by
Ωk (M ). Thus
k
! k
!
^ ^
Ωk (M ) = Γ (T ∗ M ) = Γ M, (T ∗ M ) .
(i) ω is smooth on M
(ii) M has
P an atlas such that on every chart (U, x1 , . . . , xn ), the coefficients aI of
ω = I aI dx relative to the coordinate frame {dxI }I∈Jn,k are all smooth
I
(iii) On
P everyI chart (U, x , . . . , x ) in the maximalI atlas, the coefficients aI of ω =
1 n
179
We defined 0-tensors and 0-covectors to be the constant functions, so
L0 (V ) = A0 (V ) = R.
Thus the bundle 0 (T ∗ M ) ' M × R and a 1-form on M is just a function on M . A smooth
V
0-form is thus the same as a smooth function on M . In our new notation,
0
!
ou
^
Ω0 (M ) = Γ (T ∗ M ) = Γ(M × R) = C ∞ (M ).
Similar to smooth functions, smooth differential forms can also be smoothly extended.
Zh
neighborhood of p.
We note that the extension τ̃ is not unique. It depends on p as well as the choice of a bump
function at p.
We have defined the pullback of 0-forms and 1-forms under a smooth map F : N → M . For
a smooth 0-form on M , ie a smooth function on M ,
x
F f
F ∗ f : (N −
→ M) − →R
∗
F (f ) = f ◦ F ∈ Ω0 (N ).
eli
To generalize the pullback to k-forms for k ≥ 1, we first recall the pullback of k-covectors.
A linear map L : V → W of vector spaces induces a pullback map
L∗ : Ak (W ) → Ak (V )
(L∗ α)(v1 , . . . , vk ) = α(L(v1 ), . . . , L(vk ))
for α ∈ Ak (W ) and v1 , . . . , vk ∈ V .
©F
180
for vi ∈ Tp N .
For a k-form ω on M , its pullback F ∗ ω is the k-form on N defined pointwise by
When k = 1, this formula specializes to the definition of the pullback of a 1-form. The
pullback of a k-form can be viewed as a composition
ou
F ×···×F ωF (p)
Tp N × · · · × Tp N −−∗−−−−→
∗
TF (p) M × · · · × TF (p) M −−−→ R.
Similar to the linearity of the pullback of a 0-form or 1-form, we can prove the following.
Zh
(i) F ∗ (ω + τ ) = F ∗ ω + F ∗ τ
(ii) F ∗ (aω) = aF ∗ ω
We defer the basic question of whether the pullback of a smooth k-form undder a smooth
map remains smooth for k ≥ 2.
Here vi ∈ V and σ runs over all (k, `)-shuffles of [k + `]. For 1-covectors α, β,
©F
The wedge product extends pointwise to differential forms on a manifold: for a k-form ω
and an `-form τ on M , define their wedge product ω ∧ τ to be the (k + `)-form on M such
that
(ω ∧ τ )p = ωp ∧ τp
at all p ∈ M .
181
Proposition 5.2.9
If ω, τ are smooth forms on M , then ω ∧ τ is also smooth.
ou
Zh
x
eli
©F
182
Proof
Let (U, x1 , . . . , xn ) be a chart on M . On U ,
X X
ω= aI dxI , τ= bJ dxJ
I J
ou
! !
X X
ω∧τ = aI dxI ∧ bJ dxJ
I J
X
I J
= aI bJ dx ∧ dx
I,J
!
X X
dxK .
Zh
∈ ±aI bJ
K I∪J=K,I∩J=∅
The last equality results from the observation that dxI ∧ dxJ = 0 if I, J have a common
index. If I, J are disjoint, then dxI ∧ dxJ ∈ ±dxK where K = I ∪ J but reordered as
an increasing sequence. Since the coefficients of dxK are smooth on U , we conclude the
proof.
Proof
Let (U, x1 , . . . , xn ) be P
a chart about a point
P p ∈ M and v1 , . . . , vk+` ∈ Tp N . We have the
eli
local expressions ω = I aI dxI and τ = J bJ dxJ . We have
X
= ωp (F∗,p (v1 ), . . . , F∗,p (vk ))τp (F∗,p (vk+1 ), . . . , F∗,p (vk+` ))
σ
X
= (F ∗ ω)p (v1 , . . . , vk )(F ∗ τ )p (vk+1 , . . . , vk+` )
σ
= ((F ∗ ω)p ∧ (F ∗ τ )p )(v1 , . . . , vk , vk+1 , . . . , vk+` )
= (F ∗ ω ∧ F ∗ τ )p (v1 , . . . , vk , vk+1 , . . . , vk+` ).
Define the vector space Ω∗ (M ) of smooth differential forms on a manifold M of dimension
183
n to be the direct sum n
M
∗
Ω (M ) = Ωk (M ).
k=0
This means each element Ω (M ) is uniquely a sum nk=0 ωk , where ωk ∈ Ωk (M ). With the
∗
P
wedge product, the vector space Ω∗ (M ) becomes a graded algebra, with the grading being
the degree of differential forms.
ou
5.2.7 Differential Forms on a Circle
Since the derivative ḣ(t) = (− sin, cos t) is nonzero for all t, the map h : R → S 1 is a
Zh
submersion. It can be shown in this case that the pullback by a surjective submersion is an
injective algebra homomorphism. Hence h∗ : Ω(S 1 ) → Ω∗ (R) is injective and we can identify
the differential forms on S 1 with a subspace of differential forms on R.
Let ω = −ydx + xdy be the nowhere-vanishing form on S 1 . Recall that h∗ ω = dt. Since ω is
nowhere vanishing, it is a frame for the cotangent bundle T ∗ S 1 over S 1 , and every smooth
1-form α on S 1 can be written as α = f ω for some smooth f ∈ C ∞ (S 1 ). Its pullback
f¯ := h∗ f is a smooth function on R. Since pulling back preserves multiplication,
h∗ α = (h∗ f )(h∗ ω) = f¯dt.
x
We say that a function g or a 1-form gdt on R is periodic of period a if g(t + a) = g(t) for
all t ∈ R.
Proposition 5.2.11
eli
For k = 0, 1, under the pullback map h∗ Ω∗ (S 1 ) → Ω∗ (R), smooth k-forms on S 1 are
identified with smooth periodic k-forms of period 2π on R.
Just as there are left-invariant vector fields on a Lie group G, there are also left invariant
differential forms.
`∗g (ωgx) = ωx .
184
By definition, a left-invariant k-form is uniquely determined by its value at the identity, since
for any g ∈ G,
ωg = `∗g−1 (ωe ).
Example 5.2.12
ω = −ydx + xdy is a left-invariant 1-form on S 1 .
ou
Proposition 5.2.13
Every left-invariant k-form ω on a Lie group G is smooth.
Proof
It suffices to show that for any k smooth vector fields X1 , . . . , Xk on G, the function
Zh
ω(X1 , . . . , Xk ) is smooth on G. Let (Y1 )e , . . . , (Yk )e be a basis for Te G and Y1 , . . . , Yn the
left-invariant vector fields they generate. Then Y1 , ,̇Yn is a smooth frame on G as any
left-invariant vector fields are smooth. Each Xj can be written as a linear combination
Xj = i aij Yi for some smooth functions aij . It suffices thus to show that ω(Y1 , . . . , Yk ) is
P
smooth for any left-invariant vector fields Y1 , . . . , Yk .
We have
(ω(Y1 , . . . , Yk ))(g)
= ωg ((Y1 )g , . . . (Yk )g )
= (`∗g−1 (ωe ))(`g∗ (Y1 )e , . . . , `g∗ (Yk )e ) left-invariance of both ω, Yi
x
= ωe ((Y1 )e , . . . , (Yk )e ).
k
©F
^
k
Ω (G) → G
(g∨ )
ω 7→ ωe
has an inverse defined by the left-invariant differential form generated by ωe and is therefore
an isomorphism. It follows that
k nG
dim Ω (G) = .
k
185
5.3 The Exterior Derivative
In contrast to standard calculus, the basic objects in calculus on manifolds are differential
forms rather than functions.
Recall that an antiderivation on a graded algebra A = ⊕∞
k=0 A is an R-linear map D : A → A
k
such that
ou
D(ω · τ ) = (Dω) · τ + (−1)k ω · (Dτ )
deg Dω = deg ω + m.
Zh
for all homogeneous elements ω ∈ A.
Let M be a manifold and Ω∗ (M ) the graded algebra of smooth differential forms on M . On
the graded algebra Ω∗ (M ), there is a uniquely and intrinsically defined antiderivation called
the exterior derivative. The process of applying the exterior derivative is called exterior
differentiation.
such that
(i) D is an antiderivative of degree 1
eli
(ii) D ◦ D = 0
(iii) For any f ∈ C ∞ (M ) and X ∈ (M ), (Df )(X) = Xf
Condition (iii) states that on 0-forms (functions), an exterior derivative agrees with the
differential df of a function f . Hence on a coordinate chart (U, x1 , . . . , xn ),
©F
X ∂f
Df = df = i
dxi .
i
∂x
Our goal is to prove the existence and uniqueness of an exterior derivative on a manifold.
Using its defining properties, we can then show that the exterior derivative commutes with
the pullback. As a corollary, the pullback of a smooth form by a smooth map is smooth.
186
5.3.1 Exterior Derivative on a Coordinate Chart
ou
I
for some aI ∈ C (U ).
∞
If D is an exterior derivative on U ,
D(dxI )
= D(dxi1 ∧ · · · ∧ dxik )
Zh
= (Ddxi1 ) · dxi2 ∧ · · · ∧ dxik + (−1)1 dxi1 · D(dxi2 ∧ · · · ∧ dxik )
= 0 − dxi1 · D(dxi2 ∧ · · · ∧ dxik ). D◦d=0
By an inductive argument on k, we see that DdxI = 0 for all I ∈ Jn,k , k ≥ 0. It follows that
linearity
X
Dω = D(aI dxI )
I
antiderivation
X X
= (DaI ) ∧ dxI + aI (DdxI )
I I
above
X
I
= (DaI ∧ dx )
x
I
X X ∂aI
= dxj ∧ dxI .
I j
∂xj
eli
Hence if any exterior derivative D exists, then it is uniquely defined by the expression above.
To show existence, we define D as above and show that it satisfies the three conditions. The
proof is exactly the same as in the Euclidean case and is thus omitted. We denote the unique
exterior derivative on a chart (U, φ) by dU .
©F
187
operator on W :
d
f (x) := f 0 (x).
dx
The derivative has the property that the value of f 0 (x) at a point p depends only on the
values of f in a small neighborhood of p. More precisely, if f = g on an open set U ⊆ R,
then f 0 = g 0 on U . We say that the derivative is a local operator on C ∞ (R).
ou
Definition 5.3.2 (Local Operator)
An operator D : Ω∗ (M ) → Ω∗ (M ) is said to be local if for all k ≥ 0, whenever a
k-form ω ∈ Ωk (M ) restricts to 0 on an open set U in M , then Dω ≡ 0 on U .
Zh
Example 5.3.1 (Integral Operator)
Define the integral operator
I : C ∞ [a, b] → C ∞ [a, b]
Z b
f 7→ f (t)dt.
a
We consider I(f ) as a constant function over [a, b]. This is not a local operator since I(f )
depends on the value of f over all [a, b].
x
Proposition 5.3.2
Any antiderivation D on Ω∗ (M ) is a local operator.
eli
Proof
Suppose ω ∈ Ωk (M ) and ω ≡ 0 on a open subset U . Let p ∈ U be arbitrary. We claim
that (Dω)p = 0.
Choose a smooth bump function f at p supported in U . In particular, f ≡ 1 in a
neighborhood of p within U . Then f ω ≡ 0 on M , since for a point q ∈ U , then ωq = 0
and if q ∈
/ U , we have f (q) = 0. By the antiderivation property of D to f ω,
©F
0 = D(0)
= D(f ω)
= (Df ) ∧ ω + (−1)0 ∧ f (Dω).
188
We remark that the same proof shows that a derivation on Ω∗ (M ) is also a local operator.
ou
We know that there is an exterior derivative dU on U given by
X
dU ω = daI ∧ dxI
I
on U . Define (dω)p = (dU ω)p . We need to show that (dU ω)P p is independent of the chart U
about p. If (V, y 1 , . . . , y n ) is another chart about p and ω = J bJ dy J on V , then on U ∩ V ,
Zh
X X
aI dxI = dy J .
I J
Thus (dω)p = (dU ω)p is well defined independently of the chart (U, x1 , . . . , xn ).
As p varies over all points of M , this defines an operator d : Ω∗ (M ) → Ω∗ (M ). In order
to check that d satisfies the defining properties of an exterior derivative, it suffices to check
them at each point p ∈ M , which we have already done.
©F
189
Thus Df = df on functions f ∈ Ω0 (M ).
Consider now a wedge product of exact 1-forms df 1 ∧ · · · ∧ df k :
D(df 1 ∧ · · · ∧ df k )
= D(Df 1 ∧ · · · ∧ Df k ) Df i = df i
ou
k
antiderivation
X
= (−1)i−1 Df 1 ∧ · · · ∧ DDf i ∧ · · · ∧ Df k
i=1
= 0. D2 = 0
We now show that D agrees with d on any k-form ω ∈ Ωk (M ). Fix p ∈ M and choose a
chart (U, x1 , . . . , xn ) about p so that ω = I aI dxI on U . Extend the functions aI , xi on U
P
Zh
to smooth functions ãI , x̃i on M that agree with aI , xi on a neighborhood V 3 p. Define
X
ω̃ = ãI dx̃I ∈ Ωk (M ).
I
Then ω ≡ ω̃ on V .
Since D is a local operator, Dω = Dω̃ on V . Hence
(Dω)p
x
= (Dω̃)p
X
= (D ãI dx̃I )p
I
eli
X X
=( DãI ∧ dx̃I ∧ ãI ∧ Ddx̃I )p
I I
X
=( dãI ∧ dx̃I )p Dd = DD = 0
I
X
=( daI ∧ dxI )p
I
©F
= (dω)p .
Theorem 5.3.3
On any manifold M , there exists an exterior derivative d : Ω∗ (M ) → Ω∗ (M ) charac-
terized uniquely by the three defining properties.
190
5.3.5 Exterior Differentiation Under a Pullback
We now show that the pullback of differential forms commutes with the exterior derivative.
Combined with the fact that the pullback preserves the wedge product, this is a cornerstone
of calculations involving the pullback. We use these two properties to show that the pullback
of a smooth form under a smooth form remains smooth.
ou
Proposition 5.3.4 (Commutation of the Pullback with d)
Let F : N → M be a smooth map of manifolds. If ω ∈ Ωk (M ), then dF ∗ ω = F ∗ dω.
Proof
We have already proven the case of k = 0 where ω = h ∈ C ∞ (M ) by checking that
(F ∗ dh)p (Xp ) = (dF ∗ h)p (Xp ) for any Xp ∈ Tp N . Consider now the case of k ≥ 1. We
check that dF ∗ ω = F ∗ dω at every point p ∈ N . This reduces the proof to a local
Zh
computation.
If (V, y 1 , . . . , y m ) is a chart on M about F (p), then on V
X
ω= aI dy i1 ∧ · · · ∧ dy ik
I
where I = (i1 < · · · < ik ) and aI ∈ C ∞ (V ). Since the pullback distributes across the
wedge product,
X
F ∗ω = (F ∗ aI )F ∗ dy i1 ∧ · · · ∧ F ∗ dy ik
I
x
base case
X
= (aI ◦ F )dF i1 ∧ · · · ∧ dF ik
I
X
∗
dF ω = d(aI ◦ F ) ∧ dF i1 ∧ · · · ∧ dF ik .
eli
I
= F daI ∧ F ∗ dy i1 ∧ · · · ∧ F ∗ dy ik
∗
base case
X
= d(F ∗ aI ) ∧ dF i1 ∧ · · · ∧ dF ik
I
X
= d(aI ◦ F ) ∧ dF i1 ∧ · · · ∧ dF ik .
I
By computation,
dF ∗ ω = F ∗ dω.
191
Corollary 5.3.4.1
If U ⊆ M is open and ω ∈ Ωk (M ), then
(dω)|U = d(ω|U ).
Proof
Let ι : U → M be the inclusion map. We have ω|U = ι∗ ω so that
ou
(dω)|U = ι∗ dω = dι∗ ω = d(ω|U ).
Example 5.3.5
Let U = (0, ∞) × (0, 2π) in the (r, θ)-plane R2 . Define F : U ⊆ R2 → R2 by F (r, θ) =
(r cos θ, r sin θ). Let x, y be the standard coordaintes on the target R2 . We compute
F ∗ (dx ∧ dy).
Zh
F ∗ dx = dF ∗ x
= d(x ◦ F )
= d(r cos θ)
= cos(θ)dr − r sin θdθ
F dy = dF ∗ y
∗
= d(r sin θ)
antiderivation
x
= (sin θ)dr + r cos θdθ.
Proposition 5.3.6
©F
Proof
Fix p ∈ N . We show that there is an neighborhood about p on which F ∗ ω is smooth.
Choose a chart (V, y 1 , . . . , y m ) on M about F (p). Let F i = y i ◦ F be the i-th coordi-
nate of the map F in this chart. By the continuity (smoothness) of F , there is a chart
192
(U, x1 , . . . , xn ) on N about p such that F (U ) ⊆ V . Because ω is smooth on V ,
X
ω= aI dy i1 ∧ · · · ∧ dy ik
I
ou
I
X
= (F ∗ aI )dF ∗ y i1 ∧ · · · ∧ dF ∗ y ik
I
X
= (aI ◦ F )dF i1 ∧ · · · ∧ dF ik
I
X ∂(F i1 , . . . , F ik ) J
= (aI ◦ F ) dx .
Zh
I,J
∂(xj1 , . . . , xjk )
Since aI ◦ F and ∂(F i1 , . . . , F ik )/∂(xj1 , . . . , xjk ) are all smooth, we see that F ∗ ω is smooth
as desired.
In summary, if F : N → M is a smooth map of manifolds, then the pullback map F ∗ Ω∗ (M ) →
Ω∗ (N ) is a morphism of differential graded algebras, ie a degree-preserving algebra homo-
morphism that commutes with the differential.
We remark that a nonzero form on M may restrict to the zero form on a submanifold S.
For example, if S is a smooth curve on R2 defined by the nonconstant function f (x, y), then
∂f ∂f
df = dx + dy
∂x ∂y
is a nonzero 1-form on R2 , but since f is identically zero on S, df ≡ 0 on S.
Since pullbacks and exterior differentiation commute, we can write df |S to denote either one
of (df )|S = d(f |S ).
193
5.3.7 A Nowhere-Vanishing 1-Form on the Circle
ou
manifold.
At p = (1, 0), a basis for the tangent space Tp S 1 is ∂/∂y. Although dx is a nowhere-vanishing
1-form on R2 , it vanishes at (1, 0) when restricted to S 1 as
∂
(dx)p = 0.
∂y
Zh
In order to find a nowhere-vanishing 1-form on S 1 , we take the exterior derivative of both
sides of the equation
x2 + y 2 = 1.
We get
2xdx + 2ydy = 0.
Note that this equation is valid only at a point in S 1 . Define
Ux := {(x, y) ∈ S 1 : x 6= 0}
Uy := {(x, y) ∈ S 1 : y 6= 0}.
x
By our calculations, on Ux ∩ Uy ,
dy dx
=− .
eli
x y
Define a 1-form ω on S 1 by
(
dy
x
, p ∈ Ux
ωp =
− dx
y
, p ∈ Uy .
and similarly Ux− , Uy+ , Uy− . On Ux+ , y is a local coordinate and so dy is a basis for the
cotangent space Tp∗ S 1 at each p ∈ Ux+ . Since ω = dy/x on Ux+ , ω is smooth and nowhere
zero on Ux+ . A similar argument applies to dy/x on Ux− and −dx/y on Uy+ , Uy− . Hence ω is
smooth and nowhere vanishing on S 1 .
194
5.4 The Lie Derivative & Interior Multiplication
The exterior differentiation d was first locally defined with respect to a chart. It turns out
that d is in fact global and intrinsic to the manifold. We seek to derive a global intrinsic
formula for the exterior of a k-form such as the following:
(dω)(X, Y ) = Xω(Y ) − Y ω(X) − ω([X, Y ]).
ou
The proof uses the Lie derivative and interior multiplication, two other instrinsic operations
on a manifold. The Lie derivative allows us to differentiate a vector field or a differential
form on a manifold along another vector field. For any vector field X on a manifold, the
interior multiplication ιX is an antiderivation of degree −1 on differential forms.
exists
Zh
a collection {Xt } or {ωt } of vector fields or diffferential forms on a manifold is said to be
a 1-parameter family if the parameter t runs over some subset of R. Let I ⊆ R be an
open interval and suppose {Xt } is a 1-parameter family of vector fields on M defined for all
t ∈ I \ {t0 } for some t0 ∈ I. We say that the limit
lim Xt
t→t0
a (t, p)∂/∂x |p and limt→t0 a (t, p) exists for all i. In this case, we define
x
n
X ∂
lim Xt |p := lim ai (t, p) .
t→t0
i=1
t→t0 ∂xi p
eli
It can be shown that this definition of the limit of Xt as t → t0 is independent of the choice
of the coordinate neighborhood (U, x1 , . . . , xn ), as there is a smooth change of coordinates.
A 1-parameter family {Xt }t∈I of smooth vector fields on M is said to depend smoothly on t
if every p ∈ M has a coordinate neighborhood (U, x1 , . . . , xn ) on which
X ∂
(Xt )p = ai (t, p)
©F
i
∂xi p
for (t, p) ∈ I × U and smooth functions ai on I × U . In this case we also say that {Xt }t∈I is
a smooth family of vector fields on M .
For a smooth family of vector fields on M , one can define its derivative with respect to t = t0
by !
d X ∂ai ∂
Xt = (t0 , p) i
dt t=t0 i
∂t ∂x p
p
195
for (t0 , p) ∈ I × U . It can be shown that this definition is independent of the chart
(U, x1 , . . . , xn ) containing p by considering a smooth change of coordinates. Indeed, Let
(V, y 1 , . . . , y n ) be another coordinate neighborhood of p such that
X ∂
Xt = bj (t, q)
j
∂y j
ou
on V . On the intersection U ∩ V ,
∂ X ∂y j ∂
= .
∂xi j
∂xi ∂y j
It follows that
X ∂y j
bj (t, p) = ai (t, p) .
Zh
i
∂xi
Differentiating both sides with respect to t yields
∂bj X ∂ai ∂y j
= i
.
∂t i
∂t ∂x
But then
X ∂bj ∂ X ∂ai ∂y j ∂
=
j
∂t ∂y j i,j
∂t ∂xi ∂y j
x
X ∂ai ∂
=
i
∂t ∂xi
for (t, p) ∈ I × U and some smooth functions bJ on I × U . We also call such a family {ωt }t∈I
a smooth family of k-forms on M and define its derivative with respect to t to be
!
d X ∂bJ
ωt = (t0 , p)dxJ |p .
dt t=t0 J
∂t
p
Similar to vector fields, this definition is independent of the chart and defines a smooth
k-form d/dt|t=t0 ωt on M .
196
Note that we write d/dt for the derivative of a smooth family of vector fields or differential
forms, but ∂/∂t for the partial derivative of a function of several variables.
ou
dt dt dt
Proof
Written out in local coordinates, the statement reduces to the usual product rule in
calculus.
Zh
If {ωt } is a smooth family of differential forms on a manifold M , then
d d
dωt = d ωt .
dt t=t0 dt t=t0
Proof
We first check that
d d
(dωt ) = d ωt
dt dt
at anParbitrary point p ∈ M . Indeed, let (U, x1 , . . . , xn ) be a neighborhood of p such that
x
ω = J bJ dxJ for some smooth functions bJ on I × U . On U ,
d
(dωt )
eli
dt
d X ∂bJ i
= dx ∧ dxJ
dt J,i ∂xi
X ∂ ∂bJ
= i
dxi ∧ dxJ exchange order
i,J
∂x ∂t
©F
!
X ∂bJ
=d dxJ
∂t
J
d
=d ωt .
dt
197
5.4.2 The Lie Derivative of a Vector Field
ou
at two nearby points p, q ∈ M , the tangent vectors Yp , Yq are in different vector spaces
Tp M, Tq M .so we cannot subtract them. One way around this is to use the local flow of
another vector field X to transport Yq to the tangent space Tp M at p.
Recall that for any smooth vector field X on M , there is a neighborhood U of p on which
the vector field as a local flow: ie there is some ε > 0 and a map
ϕ : (−ε, ε) × U → M
Zh
such that if we write ϕt (q) = ϕ(t, q), then
∂
ϕt (q) = Xϕt (q) ϕ0 (q) = q q ∈ U.
∂t
In other words, for each q ∈ U , the curve ϕt (q) is an integral curve of X with initial point q.
By definition, ϕ0 (q) = q. The local flow also satisfies the property
ϕs ◦ ϕt = ϕs+t
whenever both sides are defined. Thus for each t, the map ϕt : U → ϕt (U ) is a diffeomor-
phism onto its image, with the smooth inverse ϕ−t . Indeed,
x
ϕ−t ◦ ϕt = ϕ0 = Id, ϕt ◦ ϕ−t = ϕ0 = Id .
eli
Let Y be a smooth vector field on M . To compare the values of Y at ϕt (p) and at p, we use
the diffeomorphism ϕ−t : ϕt (U ) → U to push Yϕt (p) into Tp M .
tangent vector
198
In the definition above, the limit is taken in the finite-dimensional vector space Tp M . For
the derivative to exist, it suffices that {ϕ−t∗ Y } be a smooth family of vector fields on M .
To show the smoothness of the family, we write ϕ−t∗ Y in the local coordinates x1 , . . . xn in
a chart. Let ϕit and ϕi be the i-th components of ϕt , ϕ respectively. Then
(ϕt )i (p) = ϕi (t, p) = (xi ◦ ϕ)(t, p).
Recall that relative to the frame {∂/∂xj }, the differential ϕt∗ at p is represented by the
ou
Jacobian matrix
∂(ϕt )i ∂ϕi
= .
∂xj (p) ∂xj (t, p)
Hence !
∂ X ∂ϕi ∂
ϕt∗ j
= j
(t, p) i .
∂x p i
∂x ∂x ϕ t (p)
Zh
Thus if Y = j b ∂/∂x , then
P j j
!
X ∂
ϕ−t∗ (Yϕt (p) ) = bj (ϕ(t, p))ϕ−t∗
j
∂xj ϕt (p)
X ∂ϕi ∂
= bj (ϕ(t, p)) j
(−t, p) i .
i,j
∂x ∂x p
When X, Y are smooth vector fields on M , both ϕi , bj are smooth functions. Hence {ϕ−t∗ Y }
is indeed a smooth family of vector fields on M . It follows that the Lie derivative LX Y exists
and is given in local coordinates by
x
d
(LX Y )p = ϕ−t∗ (Yϕt (p) )
dt t=0
∂ϕi
X ∂
∂
eli
j
= b (ϕ(t, p)) j (−t, p) .
i,j
∂t t=0 ∂x ∂xi p
Theorem 5.4.3
If X, Y are smooth vector fields on a manifold M , then the Lie derivative LX Y
©F
199
Proof
We check the equality LX Y = [X, Y ] at every point by expanding both sides in local coor-
dinates. Suppose a local flow forPX is given by ϕ : (−ε, ε) × U → M , where (U, x1 , . . . , xn )
is a coordinate chart. Let X = i ai ∂/∂xi and Y = j bj ∂/∂xj on U . Recall that a local
P
flow c(t) of X satisfies
∂
ou
X
Xc(t) = ai (c(t))
i
∂xi c(t)
X ∂
c0 (t) = ċi (t) .
i
∂xi c(t)
Zh
i = 1, . . . , n
∂t
The initial conditions state that at t = 0,
∂ϕi
(0, p) = ai (ϕ(0, p)) = ai (p).
∂t
But ϕ(0, p) = p so ϕ0 is the identity map and its Jacobian is the identity. In particular,
∂ϕi
(0, p) = δji
∂xj
200
and the expression above simplifies above to
(LX Y )p
X ∂bj ∂ϕi
X
∂ai
∂
k j
= k
(p)a (p) j (0, p) − b (p) j (p)
i,j,k
∂x ∂x i,j
∂x ∂xi
X ∂bi i
k k ∂a ∂
= a − b
ou
i,k
∂xk ∂xk ∂xi
= [X, Y ].
Although the Lie derivative of a vector fields does not give us anything new, it is a useful
tool alongside the Lie derivative of differential forms.
Zh
5.4.3 The Lie Derivative of a Differential Form
Let X be a smooth vector field and ω a smooth k-form on a manifold M . Fix p ∈ M and let
ϕt : U → M be a flow of X in a neighborhood U of p. The definition of the Lie derivative of
a differential form is similar to that of the Lie derivative of a vector field. Instead of pushing
a tangent vector at ϕt (p) to p via (ϕ−t )∗ , we now pull the k-covector ωϕt (p) back to p via ϕ∗t .
By a similar argument to the case of vector fields, it can be shown that {ϕ∗t ω} is a smooth
©F
Proposition 5.4.4
Let f be a smooth function and X be a smooth vector field on M . Then LX f = Xf .
Proof
Fix p ∈ M and let ϕt : U → M be a local flow of X as above. Since ϕt (p) is a curve
201
through p with initial vector Xp ,
d
(LX f )p := (ϕ∗ f )p
dt t=0 t
d
:= (f ◦ ϕt )(p)
dt t=0
= Xp f.
ou
5.4.4 Interior Multiplication
Zh
We first define interior multiplication on a vector space.
for v2 , . . . , vk ∈ V .
x
We define ιv β = β(v) ∈ R for a 1-covector β on V and ιv β = 0 for a 0-covector (constant) β
on V .
Proposition 5.4.5
eli
For 1-covectors α1 , . . . , αk on a vector space V and v ∈ V ,
k
X
1 k
ιv (α ∧ · · · ∧ α ) = (−1)i−1 αi (v)α1 ∧ · · · ∧ αbi ∧ · · · ∧ αk .
i=1
Here the hat over αi indicates that αi is omitted from the wedge product.
©F
Recall that
X
(α1 ∧ · · · ∧ αk )(v1 , . . . , vk ) = (sgn σ)α1 (vσ(1) ) . . . αk (vσ(k) ) = det αi (vj ) .
σ∈Sk
202
Proof
By computation,
ou
k
expansion along 1st column
X
(−1)i+1 αi (v) det α` (vj ) `6=i,j6=2
=
i=1
Xk
= (−1)i+1 αi (v) α1 ∧ · · · ∧ αbi ∧ · · · ∧ αk (v2 , . . . , vk ).
i=1
Proposition 5.4.6
Fix a vector v in a vector space V . Let ιv : ∗ (V ∨ ) → ∗−1 (V ∨ ) be interior multiplica-
Zh
V V
tion by v. Then
(i) ιv ◦ ιv = 0
(ii) for β ∈ k (V ∨ ) and γ ∈ ` (V ∨ ),
V V
ιv (β ∧ γ) = (ιv β) ∧ γ + (−1)k β ∧ ιv γ.
Proof
x
(i) Let β ∈ k (V ∨ ). By the definition of interior multiplication,
V
β = α1 ∧ · · · ∧ αk , γ = αk+1 ∧ · · · ∧ αk+` ,
203
where the αi ’s are all 1-covectors. Then
ιv (β ∧ γ)
= ιv (α1 ∧ · · · ∧ αk+` )
k+`
X
= (−1)i−1 αi (v)α1 ∧ · · · ∧ αbi ∧ · · · ∧ αk+`
i=1
ou
k
!
X
= (−1)i−1 αi (v)α1 ∧ · · · ∧ αbi ∧ · · · ∧ αk ∧ αk+1 ∧ · · · ∧ αk+`
i=1
k
X
k 1 k
+ (−1) α ∧ · · · ∧ α ∧ (−1)i+1 αk+i (v)αk+1 ∧ · · · ∧ αd
k+i ∧ · · · ∧ αk+`
i=1
= (ιv β) ∧ γ + (−1)k β ∧ ιv γ.
Zh
Interior multiplication on a manifold is defined pointwise. If X is a smooth vector field on
M and ω ∈ Ωk (M ), then ιX ω is the (k − 1)-form defined by
for all p ∈ M . The form ιX ω on M is smooth since for any smooth vector fields X2 , . . . , Xk
on M ,
(ιX ω)(X2 , . . . , Xk ) = ω(X, X2 , . . . , Xk )
is a smooth function on M . In the case that ω is a 1-form, ιX (ω) = ω(X). If ω = f is a
-form (function) on M , then ιX f = 0. By the properties of interior multiplication at each
point p ∈ M , the map ιX : Ω∗ (M ) → Ω∗ (M ) is an antiderivation of degree −1 such that
x
ιX ◦ ιX = 0.
Let F denote the ring C ∞ (M ) of smooth functions on the manifold M . As ιX ω is a point
operator, ie its value at p depends only on Xp , ωp , it is F-linear in either argument. Thus
eli
ιX ω is additive in each argument.
Proposition 5.4.7
For any f ∈ F ,
(i) ιf X ω = f ιX ω
(ii) ιX (f ω) = f ιX ω
©F
Proof
We omit the proof of (ii) as it is similar.
204
Example 5.4.8 (Interior Multiplication on R2 )
Let X = x∂/∂x + y∂/∂y be the radial vector field and α = dx ∧ dy the area 2-form on
the plane R2 . We compute the contraction ιX α.
Firstly,
ιX dx = dx(X) = x
ou
ιX dy = y.
ιX α = ιX (dx ∧ dy)
= (ιX dx)dy − dx(ιX dy)
= xdy − ydx.
Zh
This restricts to the nowhere-vanishing 1-form ω on the circle S 1 .
We state and prove several basic properties of the Lie derivative, including its relation to
exterior derivation and interior multiplication, two other intrinsic operations on a manifold.
Theorem 5.4.9
Let X be a smooth vector field on a manifold M .
x
(i) The Lie derivative LX : Ω∗ (M ) → Ω∗ (M ) is a derivation, ie it is an R-linear
map such that for all ω ∈ Ωk (M ) and τ ∈ Ω` (M ),
eli
LX (ω ∧ τ ) = (LX ω) ∧ τ + ω ∧ (LX τ ).
k
X
LX (ω(Y1 , . . . , Yk )) = (LX ω)(Y1 , . . . , Yk ) + ω(Y1 , . . . , LX Yi , . . . , Yk ).
i=1
Recall the product rule for smooth families {ωt }, {τt } of k-forms and `-forms:
d d d
(ωt ∧ τt ) = ωt ∧ τt + ωt ∧ τt .
dt dt dt
205
Proof (i)
Let p ∈ M and ϕt : U → M a local flow of X in a neighborhood U 3 p.
The Lie derivative LX is the d/dt of a vector-valued function of t. Thus the derivation
property is really just the product rule for smooth families of differential forms:
d
(LX (ω ∧ τ ))p = (ϕ∗ (ω ∧ τ ))p
ou
dt t=0 t
d
= (ϕ∗ ω)p ∧ (ϕ∗t τ )p
dt t=0 t
= (LX ω)p ∧ τp + ωp ∧ (LX τ )p .
Recall that the exterior derivative commutes with the pullback by a smooth functions as
well with d/dt of a smooth family of differential forms.
Zh
Proof (ii)
Let p ∈ M and ϕt : U → M a local flow of X in a neighborhood U 3 p.
d
LX dω := ϕ∗ dω
dt t=0 t
d
= dϕ∗ ω
dt t=0 t
d ∗
=d ϕω
dt t=0 t
x
= dLX ω.
Recall that if A, B are both superderivations of degree m1 , m2 , then AB − (−1)m1 m2 BA is a
superderivation of degree m1 +m2 . In particular, if A, B are antiderivations (superderivations
eli
of odd degree), then AB + BA is a derivation of degree m1 + m2 .
Also recall from an earlier proposition that
LX f = Xf
for any f ∈ C (M ) and X ∈ X(X).
∞
©F
Proof (iii)
Let p ∈ M and ϕt : U → M a local flow of X in a neighborhood U 3 p. We claim that it
suffices to check that
LX f = (dιX + ιX d)f
for any f ∈ C ∞ (U ).
Indeed, for any ω ∈ Ωk (M ), it suffices to check that at any p ∈ M , LX ω = (dιX +
ιX d)ω. By shrinking U if necessary, we may assume we have a coordinate neighborhood
206
(U, x1 , . . . , xn ). Moreover, by linearity, we may further assume that ω is a wedge product
ω = f dxi1 ∧· · ·∧dxik . Next, the LHS is a derivation by (i) and the RHS is also a derivation
(superderivation of even degree). By (ii), the LHS commutes with exterior derivation and
so does the RHS:
d(dιX + ιX d) = dιX d = (dιX + ιX d)d.
thus both sides of the Cartan homotopy formula are derivations that commute with d.
ou
Thus if the formula holds for two differential forms ω, τ , then it holds for the wedge
product ω ∧ τ and dω. It follows that the reduction above is justified.
We conclude the proof by verifying for f ∈ C ∞ (U ) that
(dιX + ιX d)f = ιX df ιX f = 0
:= (df )(X)
Zh
= Xf
= LX f.
Recall that {ϕ−t∗ Y } is a smooth family of vector fields for each Y ∈ X(M ).
Also recall that ϕ−t∗ (Yϕt (p) ) is smooth and hence continuous at a neighborhood of (0, p).
Proof (iv)
Let p ∈ M and ϕt : U → M a local flow of X in a neighborhood U 3 p. The proof is
similar to that of the standard product rule for the calculus derivative and we focus on
the case of k = 2 for the sake of simplicity as the general case is similar but more tedious.
x
By the old trick of adding and subtracting terms,
ωp (ϕ−t∗ (Yϕt (p) ), ϕ−t∗ (Zϕt (p) )) − ωp (Yp , ϕ−t∗ (Zϕt (p) ))
+ lim
t→0 t
ωp (Yp , ϕ−t∗ (Zϕt (p) )) − ωp (Yp , Zp )
+ lim .
t→0 t
207
We rewrite the first limit in the summation as
(ϕ∗t ωϕt (p) )(ϕ−t∗ (Yϕt (p) ), ϕ−t∗ (Zϕt (p) )) − ωp (ϕ−t∗ (Yϕt (p) ), ϕ−t∗ (Zϕt (p) ))
t
∗
ϕ (ωϕt (p) ) − ωp
= t (ϕ−t∗ (Yϕt (p) ), ϕ−t∗ (Zϕt (p) ))
t
→ (LX ω)p (Yp , Zp ). t→0
ou
Here the limit is justified as both the operator and arguments have limits.
By the bilinearity of ωp , the second term is
Zh
Finally, by a similar calculation, the third term is given by ωp (Yp , (LX Z)p ).
Remark 5.4.10 Unlike interior multiplication, the Lie derivative LX is not F-linear in
either argument. By the derivation property,
LX (f ω) = (LX f )ω + f LX ω = (Xf )ω + f LX ω.
We note that the previous theorem can be used to compute the Lie derivative of a differential
form.
x
Example 5.4.11 (The Lie Derivative on a Circle)
Let ω be the 1-form −ydx + xdy and X the tangent vector field −y∂/∂x + x∂/∂y on the
eli
unit circle S 1 . We have
The definition of the Lie derivative only makes sense in a neighborhood of a point as it is
local. The product formula gives us access to a global formula for the Lie derivative.
208
Theorem 5.4.12 (Global Formula for the Lie Derivative)
For a smooth k-form ω and smooth vector fields X, Y1 , . . . , Yk on a manifold M ,
k
X
(LX ω)(Y1 , . . . , Yk ) = X(ω(Y1 , . . . , Yk )) − ω(Y1 , . . . , [X, Yi ], . . . , Yk ).
i=1
ou
The definition of the exterior derivative d is also local. Using the Lie derivative, we obtain
a useful global formula for the exterior derivative. We begin with the case of a 1-form.
Proposition 5.4.13
If ω is a smooth 1-form and X, Y are smooth vector fields on a manifold M , then
Zh
Proof
It suffices to check the P
formula in a chart (U, x1 , . . . , xn ), thus we assume without loss
of generality that ω = i ai dxi . Since both sides of the equation are R-linear in ω, we
further assume that ω = f dg where f, g ∈ C ∞ (U ).
In this case,
dω = d(f dg) = df ∧ dg
and
dω(X, Y ) = df (X)dg(Y ) − df (Y )dg(X) = (Xf )Y g − (Y f )Xg.
x
On the other hand,
manifold M ,
(dω)(Y0 , . . . , Yk )
k
X X
= (−1)i Yi ω(Y0 , . . . , Ybi , . . . , Yk ) + (−1)i+j ω([Yi , Yj ], Y0 , . . . , Ybi , . . . , Ybj , . . . , Yk ).
i=0 0≤i<j≤k
209
We have already shown the case of k = 1. Assuming the formula for degrees k − 1, the case
of degree k can be shown by induction. Indeed, by the Cartan homotopy formula,
The first term can be computed using the global formula for the Lie derivative LY0 ω while
ou
the second term can be computed using the induction hypothesis.
Zh
x
eli
©F
210
Chapter 6
ou
Integration
Zh
On a manifold, we integrate differential forms rather than functions. We focus on the
integration of smooth forms over a submanifold. Note that it is nontheless possible to
integrate noncontinuous forms over more general sets.
for integration over a manifold to be well-defined, the manifold must be oriented. We begin
by discussing orientations on a manifold and enlarge the category of manifolds to include
manifolds with boundary. Our treatment of integration culminates in Stokes’ theorem for
an n-dimensional manifold.
6.1 Orientations
x
Our goal is to define orientations for n-manifolds and to investigate various equivalent char-
acterizations of orientations.
eli
For this
segment, we assume all vector spaces are finite-dimensional. Two ordered bases
u = u1 . . . un and v = v1 . . . vn of a vector V are equivalent, written u ∼ v, if
©F
Note that any finite-dimensional vector space has exactly two orientations. If µ is an ori-
entation of a finite-dimensional vector space V , we denote the other orientation by −µ and
call it the opposite of the orientation µ.
211
By convention, we define an orientation on the zero-dimensional vector space to be one of
two signs +, −.
We typically write v1 , . . . , vn for a basis in a vector space. We enclose the basis (v1 , . .. , vn )
if it is an ordered basis or alternatively we write it in matrix notation v1 . . . vn . An
orientation is denoted [(v1 , . . . , vn )] where the square brackets now stand for an equivalence
class.
ou
6.1.2 Orientations & n-Covectors
Rather than an ordered basis, we can also useVnan n-covector to specify an orientation. This
approach is based on the fact that the space (V ) of n-covectors on V is one-dimensional.
∨
Zh
Lemma 6.1.1
Let u1 , . . . , un and v1 , . . . , vn be vectors in a vector space V . Suppose
n
X
uj = aij vi
i=1
β is alternating
X σ(1)
= a1 . . . aσ(n) β(vσ(1) , . . . , vσ(n) )
©F
n
σ∈Sn
β is alternating
X σ(1)
= (sgn σ)a1 . . . aσ(n)
n β(v1 , . . . , vn )
σ∈Sn
= (det A)β(v1 , . . . , vn ).
It follows immediately that as ordered bases (v1 , . . . , vn ), (u1 , . . . , un ), β(u1 , . . . , un ) and
β(v1 , . . . , vn ) have the same sign if and only if det A > 0.
We say that the n-covector β determines / specifies the orientation (v1 , . . . , vn ) if β(v1 , . . . , vn ) >
212
0. The previous lemma asserts that this is well-defined. Moreover, we see that two n-
covectors β, β 0 on V determine the same orientation if and only if β = aβ 0 for some a > 0.
We define an equivalence relation on the non-zero n-covectors on V by setting β ∼ β 0 if
they differ by a positive constant. Thus we alternatively describe an orientation of V by an
equivalence class of non-zero n-covectors.
A linear isomorphism n (V ∨ ) ' R identifies the set of non-zero n-covectors with R − {0}
V
with two connected components, each of which determines an orientation of V .
ou
Example 6.1.2
Let e1 , e2 be the standard basis for R2 and α1 , α2 its dual basis. Then the 2-covector
α1 ∧ α2 determines the counterclockwise orientation since
Zh
Example 6.1.3
Let ∂/∂x|p , ∂/∂y|p be the standard basis for Tp R2 and (dx)p , (dy)p its dual basis. Then
(dx ∧ dy)p determines the counterclockwise orientation of Tp R2 .
p ∈ U.
213
Definition 6.1.3 (Continuous Pointwise Orientation)
We say that a pointwise orientation µ on M is continuous at p ∈ M if p has a
neighborhood U on which µ is represented by a continuous frame, ie there exists
continuous vector fields Y1 , . . . , Yn on U such that µq = [(Y1,q , . . . , Yn,q )] for all q ∈ U .
ou
orientable if it has an orientation. A manifold together with an orientation is said to be
oriented.
Example 6.1.4
Rn is oriented with orientation given by the continuous global frame
(∂/∂r1 , . . . , ∂/∂rn ).
Zh
Example 6.1.5 (Open Möbius Band)
Let R denote the rectangle
The open Möbius band M is the quotient of the rectangle R by the equivalence relation
generated by (0, y) ∼ (1, −y). The interior of R is the open rectangle
Recall that a section of a tangent bundle is continuous (smooth) if and only if its coefficients
with respect to a continuous (smooth) frame are continuous (smooth) functions on U .
©F
Proof
Let µ, ν be two orientations on M . At any p ∈ M , µp , νp are orientations of Tp M . Thus
they are either the same or are opposites. Define the function f : M → {±1} by
(
1, µ p = νp ,
f (p) :=
−1, µp = −νp .
214
Fix p ∈ M . By continuity, there is a connected neighborhood U 3 p on which µ =
[(X1 , . . . , Xn )] and ν = [(Y1 , . . . , Yn )] for some continuous vector fields Xi , Yj on U . Let
A = [aij ] : U → GL(n, R) be the change of basis matrix so that
X
Yj = aij Xi .
i
The entries aij are continuous functions so that the determinant det A : U → R× is also
ou
continuous.
By the intermediate value theorem, the continuous nowhere-vanishing functions det A
on the connected set U is everywhere positive or everywhere negative, as R× has two
connected components. Hence µ = ν or µ = −v on U . Thus f is locally constant. But a
locally constant function on a connected set is constant, hence µ = ν or µ = −ν on all of
M.
Zh
6.1.4 Orientations & Differential Forms
In practice, it is easier to manipulate the nowhere-vanishing top forms that specify a point-
wise orientation. We aim to show that the continuity condition on a pointwise orientation
translates to a smooth condition on nowhere-vanishing top forms.
Lemma 6.1.7
A pointwise orientation µ = [(X1 , . . . , Xn )] on a manifold M is continuous if and only
if each p ∈ M has a coordinate neighborhood (U, x1 , . . . , xn ) on which the function
x
(dx1 ∧ · · · ∧ dxn )(X1 , . . . , Xn ) is everywhere positive.
eli
Proof
( =⇒ ) Suppose the pointwise orientation µ is continuous. By definition, every p ∈ M
has a neighborhood W on which µ is represented by a continuous frame (Y1 , . . . , Yn ).
Choose a connected coordinate neighborhood
P i (U, x , . . . , x ) of p contained in W and for
1 n
(dx1 ∧ · · · ∧ dxn )(Y1 , . . . , Yn ) = (det bij )(dx1 ∧ · · · ∧ dxn )(∂1 , . . . , ∂n ) = det bij 6= 0.
that Xj = i cj Yi has positive determinant. Applying the previous lemma once more
215
yields that on U ,
( ⇐= ) Fix p ∈ M and suppose that on its neighborhood chart (U, x1 , . . . , xn ), the function
(dx1 ∧ · · · ∧ dxn )(X1 , . . . , Xn ) > 0 over all U .
ou
By shrinking U if necessary, we have a local representation Xj = j aij ∂i . Thus
P
0 < (dx1 ∧ · · · ∧ dxn )(X1 , . . . , Xn ) = (det aij )(dx1 ∧ · · · ∧ dxn )(∂1 , . . . , ∂n ) = det aij .
Theorem 6.1.8
Zh
An n-manifold M is orientable if and only if there exists a smooth nowhere-vanishing
n-form on M .
Proof
( =⇒ ) Let [(X1 , . . . , Xn )] be an orientation on M . The previous lemma assures that each
p ∈ M has a coordinate neighborhood (U, x1 , . . . , xn ) on which
ordered basis (X1,p , . . . , Xn,p ) for Tp M such that ωp (X1,p , . . . , Xn,p ) > 0. We show that at
every point, there is a coordinate neighborhood on which (dx1 ∧ · · · ∧ dxn )(X1 , . . . , Xn ) is
everywhere positive. The previous lemma concludes the proof.
Fix p ∈ M and let (U, x1 , . . . , xn ) be a connected coordinate neighborhood of p. Then on
U , ω = f dx1 ∧ · · · ∧ dxn for a smooth nowhere-vanishing function f . Being continuous
and nowhere vanishing on a connected set, f is everywhere positive or negative on U .
By taking x̃1 = −x1 if necessary, we may assume f > 0 on U . Then on U , (dx1 ∧ · · · ∧
dxn )(X1 , . . . , Xn ) > 0 as desired.
216
It can be shown that the unit sphere S 2 ⊆ R3 is orientation. A classical theorem from
algebraic topology states that a continuous vector field on an even-dimensional sphere must
vanish somewhere. Thus although the sphere S 2 has a continuous pointwise orientation, any
global frame that represents the orientation is necessarily discontinuous.
If ω, ω 0 are nowhere-vanishing smooth n-forms on an n-manifold, then ω = f ω 0 for some
nowhere-vanishing function f on M . Locally on a chart (U, x1 , . . . , xn ), ω = hdx1 ∧ · · · ∧ dxn
and ω 0 = gdx1 ∧ · · · ∧ dxn , where h, g are smooth nowhere-vanishing functions on U . Thus
ou
f = h/g is also a smooth nowhere vanishing function on U . Since U is an arbitrary chart,
f is smooth and nowhere vanishing function on M . On a connected manifold M , such a
function is either everywhere positive or everywhere negative. Thus the nowhere-vanishing
smooth n-forms on a connected orientable manifold M are partitioned into two equivalence
classes by the equivalence relation
ω ∼ ω 0 ⇐⇒ ω = f ω 0
Zh
with f > 0.
To each orientation µ = [(X1 , . . . , Xn )] on a connected orientable manifold M , we associate
the equivalence class of smooth nowhere-vanishing n-forms ω on M such that ω(X1 , . . . , Xn ) >
0. Such an ω exists by theorem above. If µ 7→ [ω], then −µ 7→ [−ω]. On a connected ori-
entable manifold, this yields a bijective correspondance
where each side is a set of two elements. By considering one conected component at a time,
x
we see that the bijection still holds for an arbitrary orientable manifold, with each component
having two possible orientations and two equivalence classes of smooth nowhere-vanishing n-
forms. If ω is a smooth nowhere-vanishing n-form such that ω(X1 , . . . , Xn ) > 0, we say that
eli
ω determines of specifies the orientation [(X1 , . . . , Xn )] and we call ω an orientation form on
M . An oriented manifold can be described by a pair (M, [ω]), where [ω] is the equivalence
class of an orientation form on M . However, we typically just write M if the orientation is
clear from context. For example, Rn is oriented by dx1 ∧ · · · ∧ dxn unless otherwise specified.
either [−1] or [+1]. Hence a connected zero-dimensional manifold is always orientable with
its two orientations specified by ±1.
a general zero-dimensional manifold M is a countable discrete set of points and an orientatble
is given by a function that assigns to each point either 1 or −1.
217
Proposition 6.1.10
Let U, V ⊆ Rn be open, both with the standard orientation inherited from Rn . A diffeo-
morphism F : U → V is orientation-preserving if and only if the Jacobian determinant
det[∂F i /∂xj ] is everywhere positive on U .
Proof
ou
Let x1 , . . . , xn and y 1 , . . . , y n be the standard coordinates on U, V ⊆ Rn . By computation,
Zh
Thus F is orientation-preserving if and only if det[∂F i /∂xj ] is everywhere positive on U .
Proof
( =⇒ ) Let µ = [(X1 , . . . , Xn )] be an orientation on the manifold M . By a prior lemma,
©F
218
But
dy 1 ∧ · · · ∧ dy n = (det ∂y i /∂xj )dx1 ∧ · · · ∧ dxn ,
ou
entability of the atlas, det[∂y i /∂xj ] > 0, so that (∂/∂x1 |p , . . . , ∂/∂xn |p ) is equivalent to
(∂/∂y 1 |p , . . . , ∂/∂y n |p ). Thus µ is a well-defined pointwise orientation on M . Moreover,
it is continuous as every point has a coordinate neighborhood on which µ is represented
by a continuous frame.
We say two oriented atlases {(Uα , φα )} and {(Vβ , ψβ )} on a manifold M are equivalent if the
transition functions
φα ◦ ψβ−1 : ψβ (Uα ∩ Vβ ) → φα (Uα ∩ Vβ )
Zh
have positive Jacobian determinant for all α, β.
It is not difficult to show that this is an equivalence relation on the set of oriented atlases
on a manifold M . In the proof of the theorem above, an oriented atlas {(U, x1 , . . . , xn )} on
a manifold M determines an orientation p 7→ [(∂/∂x1 |p , . . . , ∂/∂xn |p )] on M . Conversely, an
orientation [(X1 , . . . , Xn )] on M gives rise to an oriented atlas {(U, x1 , . . . , xn )} on M such
that (dx1 ∧ · · · ∧ dxn )(X1 , . . . , Xn ) > 0 on U . It can be shown that the two induced maps
{equivalence classes of oriented atlases on M } ↔ {orientations on M }
are well-defined and inverse to each other. Thus we can also specify an orientation on an
x
orientable manifold by an equivalence class of oriented atlases.
For an oriented manifold M , we denote by −M the same manifold with the opposite orien-
tation. If {(U, φ)} = {(U, x1 , . . . , xn )} is an oriented atlas specifying the orientation of M ,
the an oriented atlas specifying the orientation of −M is {(U, φ̃)} = {(U, −x1 , x2 , . . . , xn )}.
eli
Hn := {(x1 , . . . , xn ) ∈ Rn : xn ≥ 0},
with the subspace topology inherited from Rn . The points x ∈ Hn with xn > 0 are called
the interior points of Hn , and the points with xn = 0 are called the boundary points of Hn .
The sets of these two points are denoted (Hn )◦ and ∂(Hn ), respectively.
If M is a manifold with boundary, then its boundary ∂M turns out to be a manifold of
dimension n − 1 where n = dim(M ◦ ). Moreover, an orientation on M induces an orientation
on ∂M .
219
6.2.1 Smooth Invariance of Domain in Rn
In order to discuss smooth functions on a manifold with boundary, we need to extend the
notion of a smooth function to allow for nonopen domains.
ou
Let ⊆ Rn be an arbitrary subset. A function f : S → Rm os smooth at p ∈ S if there
exists a neighborhood U 3 p and a smooth function f˜ : U → Rm such that f˜ = f on
U ∩ S. We say f is smooth on S if it is smooth at each point of S.
Zh
A function f : S → Rm is smooth on S ⊆ Rn if and only if there is an open set U ⊇ S
and a smooth function f˜ : U → Rm such that f = f˜|S .
Proof
Suppose f : S → Rm is indeed smooth on S ⊆ Rn . For each p ∈ S, let Up 3 p, f˜p : Up →
Rm be the neighborhood of p and smooth function on Up that restricts to f˜ on Up ∩ S.
Then {Up }p∈S is an open cover of the open submanifold
[
U := Up .
x
p∈S
Thus there is a smooth partition of unity {ϕp } subordinate to {Up }. The function f˜ :
U → R given by
eli
X
f˜ := ϕp f˜p
p∈S
in the continuous category. We use it to show that interior points and boundary points are
invariant under diffoemorphism of open subsets of Hn .
We remark that the theorem is non-trivial since a priori, we only know that f : U → S takes
220
an open subset of U (which is open in Rn ) to an open subset of S. Hence f (U ) ⊆ S is open
in S, but not necessarily in Rn .
Proof
Fix p ∈ U . Our goal is to find a neighborhood Vf (p) 3 f (p) that is open in Rn and
contained in S.
ou
Since f : U → S is a diffeomorphism, there is an open set V ⊆ Rn containing S and a
smooth map g : V → Rn such that g|S = f −1 . Thus g ◦ f = IdU : U → U is the identity
map on U . By the chain rule,
is the identity map on the tangent space Tp U . In particular, f∗,p is necessarily injective.
Since U, V have the same dimension, it follows that f∗,p : Tp U → Tf (p) V is invertible. By
Zh
the inverse function theorem, f is locally invertible at p, meaning there are open neigh-
borhoods Up 3 p in U and Vf (p) 3 f (p) in V such that f : Up → Vf (p) is a diffeomorphism.
But
Vf (p) = f (Up ) ⊆ f (U ) = S
with V ⊆ Rn open in Rn and Vf (p) ⊆ V open in V , hence Vf (p) is open in Rn as desired.
Proposition 6.2.3
Let U, V be open subsets of the upper half-space Hn and f : U → V a diffeomorphism.
Then f maps interior points to interior points and boundary points to boundary points.
x
Note here we refer to openness in the relative topology on Hn .
Proof
Let p ∈ U be an interior point of Hn . Then p is contained in an open ball B, which is
eli
open in Rn . By the smooth invariance of domain, f (B) is open in Rn as well. Thus we
necessarily have f (B) ⊆ (Hn )◦ . But then f (p) ∈ f (B) is an interior point of Hn .
If p is a boundary point in U ∩ ∂H n , then f −1 (f (p)) = p is a boundary point. But
f −1 : V → U is a diffeomorphism, and by the contrapositive of what we just proved, f (p)
cannot be an interior point.
©F
Remark 6.2.4 Replacing Euclidean spaces by manifolds throughout this section, the iden-
tical proof steps yields the smooth invariance of domain for manifolds: If there is a diffeomor-
phism between an open subset U of an n-manifold N and an arbitrary subset S of another
n-manifold M , then S must be open in M .
221
6.2.2 Manifolds with Boundary
In the upper half-space Hn , we can distinguish open sets by those disjoint from the boundary,
or those that intersect the boundary. Charts on a manifold are homeomorphim to only the
first kind of open sets. A manifold with boundry generalizes the definition of a manifold by
allowing both kinds of open sets. we say that a topological space M is locally Hn if every
point p ∈ M has a neighborhood U homeomorphim to an open subset of Hn .
ou
Definition 6.2.2 (n-Manifold with Boundary)
A topological n-manifold with boundary is a second countable, Hausdorff topological
space that is locally Hn .
Zh
subset ϕ(U ) ⊆ Hn .
In the case of n = 1, a slight modification is necessary. We need to allow two local models,
the right half-line H1 and the left half-line
L1 := {x ∈ R : x ≤ 0}.
A chart (U, ϕ) in dimension 1 consists of an open set U in M and a homeomorphism φ of U
ewith an open subset of H1 or L1 . Under this convention, if (U, x1 , . . . , xn ) is a chart of an
n-manifold with boundary, then so is (U, −x1 , x2 , . . . , xn ) for any n ≥ 1. A manifold with
boundary has dimension at least 1, since a mnaifold of dimension 0, being a discrete set of
points, necessarily has empty boundary.
x
A collection {(U, φ)} of charts is a smooth atlas if for any two charts (U, φ) and (V, ψ), the
transition map
ψ ◦ φ−1 : φ(U ∩ V ) → ψ(U ∩ V ) ⊆ Hn
eli
is s diffeomorphism. A smooth manifold with boundary is a topological manifold with bound-
ary together with a maximal smooth atlas.
A point p ∈ M is an interior point if in some chart (U, φ), the point φ(p) is an interior point
of Hn . Similarly, p is a boundary point of M if φ(p) is a boundary point of Hn . These
concepts are well-defined independent of the choice of charts by the smooth invariance of
©F
domain. Indeed, consider any other chart (V, ψ). Then ψ ◦ φ−1 sends φ(p) 7→ ψ(p), and
φ(p), ψ(p) are either both interior points or both boundary points. The set of boundary
points of M is denoted ∂M .
Most of the concepts introduced for a manifold extend word for word to a manifold with
boundary, with the only difference being that a chart can be either of two types. For example,
a function f : M → R is smooth at a boundary point p ∈ ∂M if there is a chart (U, φ) about
p such that f ◦ φ−1 is smooth at φ(p) ∈ Hn . This in turn translates to f ◦ φ−1 having a
smooth extension to a neighborhood of φ(p) ∈ Rn .
222
We may be used to other notions of interior and boundary from point-set topology, defined
for a subset A of a topological space S. A point p ∈ S is said to be an interior point of A if
there is an open subset U ⊆ S such that
p ∈ U ⊆ A.
ou
p ∈ U ⊆ S − A.
Zh
S = int(A) t ext(A) t bd(A).
In the case the subset A ⊆ S is a manifold with boundary, we call int(A) the topological
interior and bd(A) the topological boundary, to distinguish them from the manifold interior
A◦ and the manifold boundary ∂A. Nnote that the topological interior and the topological
boundary of a set depends on an ambient space, while the manifold interior and manifold
boundary are intrinsic.
D := {(x, y) ∈ H2 : y ≤ 1}.
223
6.2.3 The Boundary of a Manifold with Boundary
ou
Moreover, if (U, φ) and (V, Ψ) are two charts on M , then
ψ 0 ◦ (φ0 )−1 : φ0 (U ∩ V ∩ ∂M ) → ψ 0 (U ∩ V ∩ ∂M )
is smooth. Thus an atlas {(Uα , φα )}α for M induces an atlas {(Uα ∩ ∂M, φα |Uα ∩∂M )} for ∂M ,
making ∂M into a manifold of dimension n − 1 without boundary.
Zh
Let M be a manifold with boundary and p ∈ ∂M . As before, two smooth functions f : U → R
and g : V → R defined on neighborhoods U, V 3 p in M are said to be equivalent if they
agree on some neighborhood W ∈ p contained in U ∩ V . A germ of smooth functions at
p is an equivalence class of such functions. Along with the usual addition, multiplicaiton,
and scalar multiplication of terms, the set Cp∞ (M ) of germs of smooth functions at p is
an R-algebra. The tangent space Tp M at p is then defined to be the vector space of all
point-derivations on the algebra Cp∞ (M ).
for instance, for p ∈ ∂H2 , ∂/∂x|p and ∂/∂y|p are both derivations on Cp∞ (H2 ). The tangent
x
space Tp (H2 ) is represented by a 2-dimensional vector space with the origin at p. Since
∂/∂y|p is a tangent vector to H2 at p, its negative −∂/∂y|p is also a tangent vector at p,
although there is not curve through p in H2 with initial velocity −∂/∂y|p .
eli
As before, the cotangent space Tp∗ M is defined to be the dual of the tangent space
Tp∗ M = Hom(Tp M, R).
Differential k-forms are also as before, as sections of the vector bundleV k T ∗ M . A differential
V
k-form is smooth if it is smooth as a section of the vector bundle k T ∗ M . For example,
dx ∧ dy is a smooth 2-form on H2 .
©F
224
Example 6.2.7
The closed interval [0, 1] is a smooth manifold with boundary. With d/dx as a continuous
pointwise orientation, [0, 1] is an oriented manifold with boundary. It has an oriented at
las with two charts (U1 , φ1 ), (U2 , φ2 ) where U1 = [0, 1), φ1 (x) = x and U2 = (0, 1], φ2 (x) =
x − 1. Note that φ2 maps to L1 .
ou
6.2.5 Outward-Pointing Vector Fields
Zh
A tangent vector Tp ∈ Tp M is outward-pointing if −Xp is inward-pointing.
Example 6.2.8
On the upper half-plane H2 , the vector ∂/∂y|p is inward-pointing and the vector −∂/∂y|p
is outward-pointing at any p on the x-axis.
Proposition 6.2.9
On any manifold M with boundary ∂M , there is a smooth outward-pointing vector field
along ∂M .
Proof (Sketch)
Cover ∂M with coordinate open sets (Uα , x1α , . . . , xnα ) in M . On each Uα , the vector field
Xα = −∂/∂xnα along Uα ∩ ∂M is smooth and outward-pointing. Choose a partition of
225
unity P
{ρα }α∈A subordinate to the open cover {Uα ∩ ∂M }α∈A . Then one can check that
X := α ρα Xα is a smooth outward-pointing vector field along ∂M .
Our goal now is to show that the boundary of an orientable manifold M with boundary is
an orientable manifold without boundary. We will designate one of the orientations on the
ou
boundary as the boundary orientation. It is easily described in terms of an orientation form
or of a pointwise orientation on ∂M .
Recall that the contraction ιX ω of the k-form by X is the (k − 1)-form given by
(ιX ω)p (v2 , . . . , vk ) := ιXp ωp (v2 , . . . , vk ) := ωp (Xp , v2 , . . . , vk ).
Zh
Proposition 6.2.10
Let M be an oriented n-manifold with boundary. If ω is an orientation form on M and
X is a smooth outward-pointing vector field on ∂M , then ιX ω is a smooth nowhere-
vanishing (n − 1)-form on ∂M . Hence, ∂M is orientable.
Proof
Since ω and X are both smooth on ∂M , so is the contraction ιX ω. We argue by contra-
diction that ιX ω must be nowhere-vanishing on ∂M .
Suppose ιX ω vanishes at some p ∈ ∂M . This means that (ιX ω)p (v1 , . . . , vn−1 ) = 0 for all
x
v1 , . . . , vn−1 ∈ Tp (∂M ). Let e1 , . . . , en−1 be a basis for Tp (∂M ). Then Xp , e1 , . . . , en−1 is a
basis for Tp M as Xp ∈ / span{e1 , . . . , en−1 } and so
Proposition 6.2.11
Suppose M is an oriented n-manifold with boundary. Let p be a point of the boundary
∂M and Xp an outward-pointing tangent vector in Tp M . An ordered basis (v1 , . . . , vn−1 )
for Tp (∂M ) represents the boundary orientation at p if and only if the ordered basis
(Xp , v1 , . . . , vn−1 ) for Tp M represents the orientation on M at p.
Proof
The proof consists of unwrapping the definitions. For p ∈ ∂M , let (v1 , . . . , vn−1 ) be an
226
ordered basis for the tangent space Tp (∂M ). Then
ou
Example 6.2.12 (The Boundary Orientation on ∂Hn )
An orientation form for the standard orientation on the upper half-space Hn is ω =
dx1 ∧ · · · ∧ dxn . A smooth outward-pointing vector field on ∂H n is −∂/∂xn . By definition,
an orientation form for the boundary orientation on ∂H n is given by the contraction
Zh
= (−1)n dx1 ∧ · · · ∧ dxn−1 .
Thus the boundary orientation on ∂H1 = {0} is given by −1, the boundary orientation
on ∂H2 is given by dx1 , and the boundary orientation on ∂H3 is given by −dx1 ∧ dx2 .
Example 6.2.13
The closed interval [a, b] ⊆ R with coordinate x has a standard orientation given by the
vector field d/dx, with orientation form dx. At the right endpoint b, an outward vector
is d/dx. Hence the boundary orientation at b is given by ιd/dx (dx) = 1. Similarly, the
boundary orientation at the left endpoint a is given by ι−d/dx (dx) = −1.
x
Example 6.2.14
Suppose c : [a, b] → M is a smooth immersion whose image is a 1-dimensional manifold
C with boundary. An orientation on [a, b] induces an orientation on C via the differential
eli
c∗,p : Tp [a, b] → Tc(p) C at each p ∈ [a, b]. In a situation like this, we give C the orientation
induced from the standard orientation on [a, b]. Thus the boundary orientation on the
boundary of C is given by +1 at the endpoint c(b) and −1 at the initial point c(a).
We first recall Riemann integration for a function over a closed rectangle in Euclidean space.
By Lebesgue’s theorem, this theory can be extended to integrals over bounded subsets of Rn
whose boundary has measure zero.
The integral of an n-form with compact support in an open set of Rn is defined to be the
Riemann integral of the coefficient function. Using a partition of unity, we define the integral
of an n-form with compact support on a manifold by writing the form as a sum of forms, each
227
with compact support in a coordinate chart. We then prove the general Stokes theorem for
an oriented manifold and show how it generalizes the fundamental theorem for line integrals
as well as Green’s theorem from calculus.
ou
We assume familiarity with Riemann integration in Rn and only briefly summarize the Rie-
mann integral of a bounded function over a bounded set in Rn .
A closed rectangle in Rn is a Cartesian product R = [a1 , b1 ] × · · · × [an , bn ] of closed intervals
in R. The volume vol(R) of the closed rectangle R is defined to be
n
Y
vol(R) := (bi − ai ).
Zh
i=1
A partition of the closed interval [a, b] is a set of real numbers {p0 , . . . , pn } such that
L(f, P ) ≤ U (f, P 0 ).
©F
L(f, P ) ≤ L(f, P 0 ).
then
U (f, P 0 ) ≤ U (f, P ).
228
Any two partitions P, P 0 of the rectangle R have a common refinement Q = {Q1 , . . . , Qn }
with Qi := Pi ∪ Pi0 . It follows that
It follows that the supremum of the lower sum L(f, P ) over all partitions P of R is less than
or equal to the infimum of the upper sum U (f, P ) over all partitions P of R. We define these
ou
two numbers to the lower integral f and the upper integral R f , respectively:
R R
R
Z Z
f := sup L(f, P ), f := inf L(f, P ).
P R P
R
Definition 6.3.1
Zh
Let R be a closed rectangle in Rn . A bounded function f : R → R is said to be
Riemann integrable if Z Z
f= f.
R R
Z Z
1 n
f (x)dx . . . dx = f˜(x)dx1 . . . dxn
A R
if the RHS exists. In this way, we can deal with the integral of a bounded function whose
domain is an arbitrary bounded set in Rn .
The volume vol(A) of a subset A ⊆ Rn is defined to be the integral A 1dx1 . . . dxn provided
R
the integral exists. This concept generalizes the volume of a closed rectangle we previously
defined.
229
6.3.2 Integrability Conditions
In this section, we describe some conditions under which a function defined on an open
subset Rn is Riemann integrable.
Recall that a set A ⊆ Rn is said to have (Lebesgue) measure zero P
if for every ε > 0, there is
a countable cover {Ri }i≥1 of A by closed rectangles Ri such that i≥1 vol(Ri ) < ε.
ou
The most useful criterion of Riemann integrability is due to Lebesgue:
Zh
Proposition 6.3.3
If a continuous function f : U → R defined on an open subset U ⊆ Rn has compact
support, then f is Riemann integrable on U .
Proof
Being continuous on a compact set, the function f is bounded. Being compact, the set
supp f is closed and bounded in Rn . We claim that the extension f˜ is continuous.
Since f˜ agrees with f on U , the extended function f˜ is continuous on U . It remains to
show f˜ is continuous on the complement. If p ∈ / U , then p ∈
/ supp f . Since supp f is a
closed subset of R , there is an open ball B 3 p disjoint from supp f . ≡f
˜ on B implying
x
n
that f˜ is continuous at p ∈
/ U . By Lebesgue’s theorem, f is Riemann integrable on U .
Remark 6.3.4 The support of a real-valued function is the closure in its domain of the
subset where the function is not zero.
eli
Definition 6.3.2 (Domain of Integration)
A subset A ⊆ Rn is called a domain of integration if it is bounded and its topological
boundary bd(A) is a set of measure zero.
Familiar figures such as triangles, rectangles, and circular disks are all domains of integration
©F
in R2 .
Proposition 6.3.5
Every bounded continuous function f defined on a domain of integration A ⊆ Rn is
Riemann integrable over A.
Proof
Let f˜ : Rn → R be the extension of f by zero. Since f is continuous on A, the extension
230
is necessarily continuous at all interior points of A. But every exterior point has a neigh-
borhood disjoint from A on which f˜ ≡ 0. But then the set of discontinuities is contained
in bd(A), which has measure zero. We conclude the proof by Lebesgue’s theorem.
ou
Once a set of coordinates x1 , . . . , xn has been fixed on Rn , n-forms on Rn can be identified
with functions on Rn , since every n-form on Rn can be written as ω = f (x)dx1 ∧ · · · ∧ dxn for
a unique function f (x) on Rn . In this way, the theory of Riemann integration of functions
on Rn carries over to n-forms on Rn .
Zh
standard coordinates x1 , . . . , xn . Its integral over a subset A ⊆ U is defined to be the
Riemann integral of f (x):
Z Z Z
1 n
ω= f (x)dx ∧ · · · ∧ dx := f (x)dx1 . . . dxn ,
A A A
Note that in this definition, we require the n-form to be written in the order dx1 ∧ · · · ∧ dxn .
If it is in any other order, we would need to rearrange it by the alternating property.
x
Example 6.3.6
If f is a bounded continuous function defined on a domain of integration A ⊆ Rn , then
the integral A f dx1 ∧ · · · ∧ dxn exists.
R
eli
Let us see how the integral of an n-form ω = f dx1 ∧ · · · ∧ dxn on an open subset U ⊆ Rn
transofrm under a change of variables. A change of variables on U is given by a diffeo-
morphism T : V ⊆ Rn → U ⊆ Rn . Let x1 , . . . , xn be the standard coordinates on U and
y 1 , . . . , y n the standard coordinates on V . Then T i := xi ◦ T = T ∗ (xi ) is the i-th component
of T . We assume that U, V are connected, and write x = (x1 , . . . , xn ) and y = (y 1 , . . . , y n ).
Furthermore, denote by J(T ) the Jacobian matrix [∂T i /∂y j ]. By the local formula for a
©F
wedge of differentials,
dT 1 ∧ · · · ∧ dT n = det(J(T ))dy 1 ∧ · · · ∧ dy n .
Recall also that the wedge product commutes with the wedge product:
F ∗ (ω ∧ τ ) = F ∗ ω ∧ F ∗ τ.
231
Hence
Z Z
∗
T ω= (T ∗ f )T ∗ dx1 ∧ · · · ∧ T ∗ dxn
V ZV
= (f ◦ T )dT 1 ∧ · · · ∧ dT n T ∗ d = dT ∗
ZV
(f ◦ T ) det(J(T ))dy 1 ∧ · · · ∧ dy n
ou
=
ZV
= (f ◦ T ) det(J(T ))dy 1 . . . dy n .
V
Recall that the change of variables formula from calculus (whose most intuitive proof is
obtained from a measure-theoretic argument regarding the Lebesgue integral) gives
Z Z
Zh
f dx . . . dx = (f ◦ T )|det J(T )|dy 1 . . . dy n
1 n
U V
232
n-form with compact support on the open subset φ(U ) ⊆ Rn . We define the integral of ω
on U to be Z Z
ω := (φ−1 )∗ ω.
U φ(U )
If (U, ψ) is another chart in the oriented atlast with the same U , then φ ◦ ψ −1 : ψ(U ) → φ(U )
is by definition an orientation-preserving diffeomorphism, and so
ou
Z Z Z
−1 ∗ −1 ∗ −1 ∗
(φ ) ω = (φ ◦ ψ ) (φ ) ω = (ψ −1 )∗ ω.
φ(U ) ψ(U ) ψ(U )
Thus the integral U ω on a chart U of the atlas is well-defined, independent of the choice of
R
Zh
Now let ω ∈ Ωnc (M ). Choose a partition of unity {ρα } subordinate to the open cover {Uα }.
Because ω has compact support and a partition of unity has locally finite supports, all except
finitely many ρα ω are identically zero. In particular,
X
ω= ρα ω
α
is a finite sum. Recall the elementary topological fact that A ∩ B ⊆ A ∩ B. this means that
supp(ρα ω) ⊆ supp(ρα ) ∩ supp(ω).
In particular, supp(ρα ω) is a closed subset of the compact set supp ω and Ris hence compact.
x
Since ρα ω is an n-form with compact support in the chart Uα , its integral Uα ρα ω is defined.
Thus we can define the integral of ω over M to be the finite sum
Z XZ
ω := ρα ω.
eli
M α Uα
For this integral to be well-defined, we must show that it is independent of the choices of
oriented atlas and partition of unity. Let {Vβ } be another oriented atlas of M specifying
the same orientation and {χβ } a partition of unity subordinate to {Vβ }. Then {(Uα ∩
Vβ , φα |Uα ∩Vβ )} and {(Uα ∩ Bβ , ψ|Uα ∩Vβ )} are two new atlases of M specifying the orientation
of M , and
©F
XZ XZ X X
ρα ω = ρα χβ ω χβ = 1
α Uα α Uα β β
XXZ
= ρα χβ ω finite sums
α β Uα
XXZ
= ρα χβ ω,
α β Uα ∩Vβ
233
where the last line follows since supp(ρα χβ ) ⊆ Uα ∩ Vβ . By symmetry, χβ ω is equal
P R
β Vβ
to the same sum. Hence
XZ XZ
ω= χβ ω,
α Uα β Vβ
ou
as desired.
Proposition 6.3.7
Let ω be an n-form with compact support on an oriented n-manifold. If −M denotes
the same manifold with the oppposite orientation, then
Z Z
Zh
ω=− ω.
−M M
Proof
By the definition of an integral, it suffices to show that for every chart (U, φ) = (U, x1 , . . . , xn )
and differential form τ ∈ Ωnc (U ), if (U, φ̄) = (U, −x1 , x2 , . . . , xn ) is the chart with the op-
posite orientation, then Z Z
−1 ∗
(φ̄ ) τ = − (φ−1 )∗ τ.
φ̄(U ) φ(U )
Similarly,
©F
234
It follows that
Z
(φ̄−1 )∗ τ
φ̄(U )
Z
=− (f ◦ φ̄−1 )dr1 . . . drn calculations above
Zφ̄(U )
(f ◦ φ−1 ) ◦ (φ ◦ φ̄−1 )|J(φ ◦ φ̄−1 )|dr1 . . . drn
ou
=−
φ̄(U )
Z
=− (f ◦ φ−1 )dr1 . . . drn change of variables in Rn
φ(U )
Z
=− (φ−1 )∗ τ.
φ(U )
Our treatment of integration above can be extended nearly verbatim to oriented manifolds
Zh
with boundary. It has the virtue of simplicity and utility in proving theorems. However, it
is not practical for the actual computation of integrals. It is best to consider integrals over
a parameterized set for explicit integral calculations.
that A is a manifold, it agrees with the earlier definition of integration over a manifold.
Subdividing an oriented manifold into a union of parameterized sets can be an effective
method of calculating an integral over the manifold.
©F
origin, ϕ is the angle that the vector hx, y, zi makes with the positive z-axis, and θ is the
angle that the vector hx, yi in the (x, y)-plane makes with the positive x-axis. Let ω be
235
the 2-form on the unit sphere S 2 ⊆ R3 given by
dy∧dz
x , f 6= 0,
ω = dz∧dx
y
, y= 6 0,
dx∧dy
z
, z 6= 0.
We with to compute
R
ω.
ou
S2
respect to the Euclidean metric. Therefore, the integral S 2 ω is the surface area of the
sphere.
The sphere S 2 has a parametrization by spherical coordinates
Zh
on D := {(ϕ, θ) ∈ R62 : ϕ ∈ [0, π], θ ∈ [0, 2π]}. Since
F ∗ x = sin ϕ cos θ,
F ∗ y = sin ϕ sin θ,
F ∗ z = cos ϕ,
we have
F ∗ dy = dF ∗ y = cos ϕ sin θdϕ + sin ϕ cos θdθ
and
x
F ∗ dz = − sin ϕdϕ,
so for x 6= 0,
F ∗ dy ∧ F ∗ dz
F ∗ω = = sin ϕdϕ ∧ dθ.
eli
F ∗x
For y 6= 0 and z 6= 0, similar calculations show that F ∗ ω is given by the same formula.
Therefore, F ∗ ω = sin ϕdϕ ∧ dθ everywhere on D, and
Z Z
ω= F ∗ω
S 2
ZD2π Z π
©F
= sin ϕdϕdθ
0 0
= 2π [− cos ϕ]π0
= 4π.
236
6.3.5 Integration over a Zero-Dimensional Manifold
The discussion of integration so far implicitly assumes that the manifold M has dimension
n ≥ 1. We now treat integration over a zero-dimensional manifold. A compact oriented
0-manifold
P M is P a finite collection of points, each point oriented by +1, −1. We write this
as M = i pi − j qj . The integral of a 0-form f : M → R is defined to be the sum
ou
Z X X
f := f (pi ) − f (qj ).
M i j
Zh
the boundary orientation and let
R ι : ∂M → M beR the ∗inclusion map. If ω is an (n − 1)-form
on M , it is customary to write ∂M ω instead of ∂M ι ω.
Proof
x
Choose an atlas {(Uα , φα )} for m in which each Uα is diffeormorphic to either Rn or
Hn via an orientation-preserving diffeomorphism. This is possible since any open disk is
diffeomorphic to Rn and any half-disk containing its boundary diameter is diffeomorphic
eli
to Hn . Let {ρα } be a smooth partition of unity subordinate to {Uα }. We showed in the
preceding section that ρα ω has compact support in Uα .
Suppose Stokes’ theorem holds for Rn and Hn . Then it holds for all the charts in our
atlas, which are diffeomorphic to Rn or Hn . This is because we defined
Z Z
(φ−1 ∗
©F
dω := α ) ω
Uα φα (U )
with φα (U ) = Rn or Hn . Similarly,
Z Z Z
∗
ω := ι ω := (φ−1 ∗
α ) ω
∂Uα ∂Uα φα (Uα )
237
Therefore,
Z Z X X
ω= ρα ω ρα = 1
∂M ∂M α α
XZ
= ρα ω finite sum
α ∂M
XZ
ou
= ρα ω supp(ρα ω) ⊆ Uα
α ∂Uα
XZ
= d(ρα ω) Stokes’ for Uα
α Uα
XZ
= d(ρα ω) supp d(ρα ω) ⊆ supp(ρα ω) ⊆ Uα
α M
Zh
Z !
finite sum
X
= d ρα ω
M α
Z
= dω.
M
Thus it suffices to prove Stokes’ theorem for Rn and Hn . We give a proof for H2 for the
sake of simplicity.
Proof of Stokes’ theorem in H2 : Let x, y be coordinates on H2 . Then the standard ori-
entation on H2 is given by dx ∧ dy, and the boundary orientation on ∂H2 is given by
x
ι−∂/∂y (dx ∧ dy) = dx.
The form ω is a linear combination
eli
ω = f (x, y)dx + g(x, y)dy
for smooth functions f, g with compact support in H2 . Since the supports of f, g are
compact, we may choose a real number a > 0 sufficiently large so that the supports of
f, g are contained in the interior of the square [−a, a] × [0, a]. We write fx , fy to denote
the partial derivatives of f with respect to x, y, respectively. Then
©F
∂g ∂f
dω = − dx ∧ dy = (gx − fy )dx ∧ dy,
∂x ∂y
and
Z Z Z
dω = gx dxdy − fy dxdy
H2 H2 H2
Z aZ a Z aZ a
= gx dxdy − fy dydx.
0 −a −a 0
238
In the expression above,
Z a a
gx (x, y)dx = g(x, y) =0
−a x=−a
ou
fy (x, y)dy = f (x, y) = −f (x, 0)
0 y=0
Zh
On the other hand, ∂H2 is the x-axis and dy = 0 on ∂H2 . It follows from the definition
of ω that Z Z a
ω= f (x, 0)dx.
∂H2 −a
We now apply Stokes’ theorem for manifolds to unify theorems of vector calculus on R2 , R3 .
Recall the calculus notation F ·dr = P dx+Qdy+Rdz for a function vector field F = hP, Q, Ri
and coordinates r = (x, y, z). As in calculus, we assume in this section that functions, vector
x
fields, and regions of integration have sufficient smoothness or regularity properties so that
all the integrals are defined.
eli
Theorem 6.3.10 (Fundamental Theorem for Line Integrals)
Let C be a curve in R3 , parameterized by some r(t) = (x(t), y(t), z(t)), t ∈ [a, b] and
let F be a vector field on R3 . If F = grad f for some scalar function f , then
Z
F · dr = f (r(b)) − f (r(a)).
C
©F
Suppose in Stokes’ theorem we take M to be a curve with parameterization r(t), t ∈ [a, b],
and ω to be the function f on C. Then
Z Z Z Z
∂f ∂f ∂f
dω = df = dx + dy + dz = grad f · dr
C C C ∂x ∂y ∂z C
and Z r(b)
ω=f = f (r(b)) − f (r(a)).
∂C r(a)
239
In this case Stokes’ theorem specializes to the fundamental theorem for line integrals.
ou
∂D D
To obtain Green’s theorem, let M be a plane region D with boundary ∂D and let ω be the
1-form P dx + Qdy on D. Then
Z Z
ω= P dx + Qdy
∂D ∂D
and
Zh
Z Z
dω = Py dy ∧ dx + Qx dy ∧ dy
D ZD
= (Qx − Py )dx ∧ dy
ZD
= (Qx − Py )dxdy
ZD
= (Qx − Py )dxdy.
D
x
In this case, Stokes’ theorem specializes to Green’s theorem in the plane.
eli
©F
240