Differentiable Manifolds: Lecture Notes For Geometry 2
Differentiable Manifolds: Lecture Notes For Geometry 2
Henrik Schlichtkrull
Department of Mathematics
University of Copenhagen
i
Preface
The purpose of these notes is to introduce and study differentiable mani-
folds. Differentiable manifolds are the central objects in differential geometry,
and they generalize to higher dimensions the curves and surfaces known from
Geometry 1. Together with the manifolds, important associated objects are
introduced, such as tangent spaces and smooth maps. Finally the theory
of differentiation and integration is developed on manifolds, leading up to
Stokes’ theorem, which is the generalization to manifolds of the fundamental
theorem of calculus.
These notes continue the notes for Geometry 1, about curves and surfaces.
As in those notes, the figures are made with Anders Thorup’s spline macros.
The notes are adapted to the structure of a course which stretches over 9
weeks. There are 9 chapters, each of a size that it should be possible to cover
in one week. For a shorter course one can read the first 6 chapters. The notes
were used for the first time in 2006. The present version has been revised,
and I would like to thank Andreas Aaserud, Hans Plesner Jakobsen and Kang
Li for many corrections and suggestions of improvements. Further revision
is undoubtedly needed, and comments and corrections will be appreciated.
Henrik Schlichtkrull
October, 2011
ii
Contents
1. Manifolds in Euclidean space. . . . . . . . . . . . . . . . . 1
1.1 Parametrized manifolds . . . . . . . . . . . . . . . . . 1
1.2 Embedded parametrizations . . . . . . . . . . . . . . . 3
1.3 Curves . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Surfaces . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Chart and atlas . . . . . . . . . . . . . . . . . . . . 8
1.6 Manifolds . . . . . . . . . . . . . . . . . . . . . . . 10
1.7 The coordinate map of a chart . . . . . . . . . . . . . . 11
1.8 Transition maps . . . . . . . . . . . . . . . . . . . . 13
1.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . 15
2. Abstract manifolds . . . . . . . . . . . . . . . . . . . . . 17
2.1 Topological spaces . . . . . . . . . . . . . . . . . . . 17
2.2 Abstract manifolds . . . . . . . . . . . . . . . . . . . 19
2.3 Examples . . . . . . . . . . . . . . . . . . . . . . . 20
2.4 Projective space . . . . . . . . . . . . . . . . . . . . 21
2.5 Product manifolds . . . . . . . . . . . . . . . . . . . 22
2.6 Smooth functions on a manifold . . . . . . . . . . . . . 23
2.7 Smooth maps between manifolds . . . . . . . . . . . . . 25
2.8 Lie groups . . . . . . . . . . . . . . . . . . . . . . . 28
2.9 Countable atlas . . . . . . . . . . . . . . . . . . . . 29
2.10 Whitney’s theorem . . . . . . . . . . . . . . . . . . . 30
2.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . 31
3. The tangent space . . . . . . . . . . . . . . . . . . . . . 33
3.1 The tangent space of a parametrized manifold . . . . . . . 33
3.2 The tangent space of a manifold in Rn . . . . . . . . . . 34
3.3 The abstract tangent space . . . . . . . . . . . . . . . 35
3.4 Equivalence of curves . . . . . . . . . . . . . . . . . . 37
3.5 The vector space structure . . . . . . . . . . . . . . . 38
3.6 Directional derivatives . . . . . . . . . . . . . . . . . 39
3.7 Action on functions . . . . . . . . . . . . . . . . . . . 40
3.8 The differential of a smooth map . . . . . . . . . . . . . 41
3.9 The standard basis . . . . . . . . . . . . . . . . . . . 45
3.10 Orientation . . . . . . . . . . . . . . . . . . . . . . 45
3.11 Exercises . . . . . . . . . . . . . . . . . . . . . . . 46
4. Submanifolds . . . . . . . . . . . . . . . . . . . . . . . 49
4.1 Submanifolds in Rk . . . . . . . . . . . . . . . . . . . 49
4.2 Abstract submanifolds . . . . . . . . . . . . . . . . . 49
4.3 The local structure of submanifolds . . . . . . . . . . . . 51
4.4 Level sets . . . . . . . . . . . . . . . . . . . . . . . 55
4.5 The orthogonal group . . . . . . . . . . . . . . . . . . 57
4.6 Domains with smooth boundary . . . . . . . . . . . . . 58
iii
C = γ(I)
W
γ(V )
x
1.3 Curves
As mentioned in the introduction, we shall define a concept of manifolds
which applies to subsets of Rn rather than to parametrizations. In order
to understand the definition properly, we begin by the case of curves in R2 .
The idea is that a subset of R2 is a curve, if in a neighborhood of each of its
points it is the image of an embedded parametrized curve.
C p C ∩ W = γ(I)
W
x
C p W
γ(I) = C ∩ W
Example 1.3.4. An 8-shaped set like the one in Example 1.2.2 is not
a curve in R2 . In that example we showed that the parametrization by
(cos t, cos t sin t) was not embedded, but of course this does not rule out that
6 Chapter 1
is a curve in R2 .
Proof. By continuity of the partial derivatives, the set of non-critical points in
Ω is an open subset. If we replace Ω by this set, the set C can be expressed as
a level curve {p ∈ Ω | f (p) = c}, to which we can apply the implicit function
theorem (see Geometry 1, Corollary 1.5). It then follows from Lemma 1.3
that C is a curve.
1.4 Surfaces
We proceed in the same fashion as for curves.
Definition 1.4. A surface in R3 is a non-empty set S ⊂ R3 satisfying the
following property for each point p ∈ S. There exists an open neighborhood
W ⊂ R3 of p and an embedded parametrized surface σ: U → R3 with image
σ(U ) = S ∩ W. (1.3)
z
W
S
σ(U ) = S ∩ W
is a surface in R3 .
Proof. The proof, which combines Geometry 1, Corollary 1.6, with Lemma
1.4 below, is entirely similar to that of Theorem 1.3.
Example 1.4.3. Let us verify for the sphere that it contains no critical
points for the function f (x, y, z) = x2 + y 2 + z 2 . The partial derivatives
are fx = 2x, fy = 2y, fz = 2z, and they vanish simultaneously only at
(x, y, z) = (0, 0, 0). This point does not belong to the sphere, hence it is a
surface. The verification for the cylinder is similar.
Lemma 1.4. Let S ⊂ R3 be non-empty. Then S is a surface if and only if
it satisfies the following condition for each p ∈ S:
There exist an open neighborhood W ⊂ R3 of p, such that S ∩ W is
the graph of a smooth function h, where one of the variables x1 , x2 , x3 is
considered a function of the other two variables.
Proof. The proof is entirely similar to that of Lemma 1.3.
where x = −1, and σ2 covers with the exception of a vertical line on the front
where x = 1. Together they cover the entire set and thus they constitute an
atlas.
z
y
x
are charts on the unit sphere. The restrictions on u and v ensure that they
are regular and injective. The chart σ covers the sphere except a half circle
(a meridian) in the xz-plane, on the back where x ≤ 0, and the chart σ̃
similarly covers with the exception of a half circle in the xy-plane, on the
front where x ≥ 0 (half of the ‘equator’). As seen in the following figure the
excepted half-circles are disjoint. Hence the two charts together cover the
full sphere and they constitute an atlas.
z
y
x
1.6 Manifolds
We now return to the general situation where m and n are arbitrary
integers with 0 ≤ m ≤ n.
Definition 1.6.1. An m-dimensional manifold in Rn is a non-empty set
S ⊂ Rn satisfying the following property for each point p ∈ S. There exists
an open neighborhood W ⊂ Rn of p and an m-dimensional embedded (see
Definition 1.2.2) parametrized manifold σ: U → Rn with image σ(U ) = S∩W.
The surrounding space Rn is said to be the ambient space of the manifold.
Clearly this generalizes Definitions 1.3 and 1.4, a curve is a 1-dimensional
manifold in R2 and a surface is a 2-dimensional manifold in R3 .
Example 1.6.1 The case m = 0. It was explained in Section 1.1 that a
0-dimensional parametrized manifold is a map R0 = {0} → Rn , whose image
consists of a single point p. An element p in a set S ⊂ Rn is called isolated
if it is the only point from S in some neighborhood of p, and the set S is
called discrete if all its points are isolated. By going over Definition 1.6.1 for
the case m = 0 it is seen that a 0-dimensional manifold in Rn is the same as
a discrete subset.
Example 1.6.2 If we identify Rm with {(x1 , . . . , xm , 0 . . . , 0)} ⊂ Rn , it is
an m-dimensional manifold in Rn .
Example 1.6.3 An open set Ω ⊂ Rn is an n-dimensional manifold in Rn .
Indeed, we can take W = Ω and σ = the identity map in Definition 1.6.1.
Example 1.6.4 Let S ⊂ S be a relatively open subset of an m-dimensional
manifold in Rn . Then S is an m-dimensional manifold in Rn .
The following lemma generalizes Lemmas 1.3 and 1.4.
Lemma 1.6. Let S ⊂ Rn be non-empty. Then S is an m-dimensional
manifold if and only if it satisfies the following condition for each p ∈ S:
There exist an open neighborhood W ⊂ Rn of p, such that S ∩ W is the
graph of a smooth function h, where n − m of the variables x1 , . . . , xn are
considered as functions of the remaining m variables.
Proof. The proof is entirely similar to that of Lemma 1.3.
Theorem 1.6. Let f : Ω → Rk be a smooth function, where k ≤ n and where
Ω ⊂ Rn is open, and let c ∈ Rk . If it is not empty, the set
is an n−k-dimensional manifold in Rn .
Proof. Similar to that of Theorem 1.3 for curves, by means of the implicit
function theorem (Geometry 1, Corollary 1.6) and Lemma 1.6.
Manifolds in Euclidean space 11
Example 1.6.5 In analogy with Example 1.4.3 we can verify that the m-
sphere
S m = {x ∈ Rm+1 | x21 + · · · + x2m+1 = 1}
then
2x1 2x2 0 0
Df (x) =
0 0 2x3 2x4
and it is easily seen that this matrix has rank 2 for all x ∈ S.
σ(V ) = S ∩ W, (1.5)
σ̃(π(p)) = p
σ̃ ◦ (π ◦ σ) = σ. (1.7)
By the chain rule for smooth maps we have the matrix product equality
D(σ̃)D(π ◦ σ) = Dσ,
and since the n × m matrix Dσ on the right has independent columns, the
determinant of the m × m matrix D(π ◦ σ) must be non-zero in each x ∈ U1
(according to the rule from matrix algebra that rank(AB) ≤ rank(B)).
By the inverse function theorem, there exists open sets V ⊂ U1 and Ṽ ⊂ Ũ
around q0 and π(σ(q0 )), respectively, such that π ◦ σ restricts to a diffeomor-
phism of V onto Ṽ . Note that σ(V ) = σ̃(Ṽ ) by (1.7).
Manifolds in Euclidean space 13
σ(U ) S
σ̃(Ũ )
σ σ̃ π
π◦σ
V Ṽ
U Ũ
the transition map between the charts. We will show that such a change of
coordinates is just a reparametrization.
Let Ω ⊂ Rk be open and let f : Ω → Rn be a smooth map with f (Ω) ⊂ S.
Let σ: U → Rn be a chart on S, then the map σ −1 ◦ f , which is defined on
σ2−1 ◦ σ1 : V1 → V2 ⊂ Rm
is a diffeomorphism.
σ2 (U )
σ1 (U )
S ⊂ Rn
σ1 σ2
U1 ⊂ R m U2 ⊂ R m
V1 σ2−1 ◦ σ1 V2
1.9 Exercises
t t2
γ(t) = (− , − )
1 + t3 1 + t3
C = {(x, y) ∈ R2 | x3 + xy + y 3 = 0}.
C = {(x, y) ∈ R2 | x3 + xy + y 3 = c}.
1
Verify that C is a curve in R2 if c
= 27 .
1
6 Let c = 27
and let C be as above. Prove or disprove that it is a curve in
R2 .
7 Prove Lemma 1.4.
8 Find an atlas for the cylinder (Example 1.5.2), consisting of a single chart.
9 Let S = {(x, y, z) | z = 0} be the xy-plane, let U = {(u, v) ∈ R2 | u > v},
and let σ : U → R3 be given by σ(u, v) = (u + v, uv, 0). Verify that σ
is a chart on S. After that, replace the condition u > v by u
= v in the
definition of U , and prove that σ is no longer a chart.
10 Give an example of a level set S = {(x, y, z) | f (x, y, z) = 0} which is a
surface in R3 , and yet there exists a point in S for which ( ∂f , ∂f , ∂f ) =
∂x ∂y ∂z
(0, 0, 0).
11 Let S = {(x, y, z) | xyz = c} where c ∈ R. Show that S is a surface in
R3 if c
= 0. Determine S explicitly when c = 0 and prove that it is not a
surface.
16 Chapter 1
x2 + y 2 + z 2 = 1 and x2 = yz 2 .
S = {(x, y, z) | ( x2 + y 2 , z) ∈ C}.
Abstract manifolds
2.3 Examples
Example 2.3.1 Let M be an m-dimensional real vector space. Fix a basis
v1 , . . . , vm for M , then the map
σ: (x1 , . . . , xm ) → x1 v1 + · · · + xm vm
τ : (x1 , . . . , xm ) → x1 w1 + · · · + xm wm
Abstract manifolds 21
x2 x3
σ1−1 (p) = ( , )
x1 x1
when p = π(x). The ratios xx21 and xx31 are continuous functions on R3 \ {x1 =
0}, hence σ1−1 ◦ π is continuous.
2) The overlap between σi and σj satisfies smooth transition. For example
1 u2
σ1−1 ◦ σ2 (u) = ( , ),
u1 u1
which is smooth R2 \ {u | u1 = 0} → R2 .
new notion of smoothness agrees with the previous one when f is defined
on an open set X in Rn . We also observe that the notion of smoothness
is independent of the target set Y , that is, f is smooth into Y if and only
if it is smooth, considered just as a map into Rl . Finally, we note that if
p ∈ X ⊂ X ⊂ Rn and f : X → Rl is smooth at p, then the restriction f |X is
also smooth at p.
Definition 2.6.2. Let S ⊂ Rn and S̃ ⊂ Rl be manifolds. A map f : S → S̃
is called smooth if it is smooth according to Definition 2.6.1 with X = S and
Y = S̃.
In particular, the above definition can be applied with S̃ = R. A smooth
map f : S → R is said to be a smooth function, and the set of these is denoted
C ∞ (S). It is easily seen that C ∞ (S) is a vector space when equipped with
the standard addition and scalar multiplication of functions. We remark that
if F ∈ C ∞ (W ) where W ⊂ Rn is open, then f = F |S ∈ C ∞ (S) if S ⊂ W .
Example 2.6.1 The functions x → xi where i = 1, . . . , n are smooth func-
tions on Rn . Hence they restrict to smooth functions on every manifold
S ⊂ Rn .
Example 2.6.2 Let S ⊂ R2 be the circle {x | x21 + x22 = 1}, and let
Ω = S \ {(−1, 0)}. The function f : Ω → R defined by f (x1 , x2 ) = 1+x x2
1
is a
smooth function, since it is the restriction of the smooth function F : W → R
defined by the same expression for x ∈ W = {x ∈ R2 | x1
= −1}.
Example 2.6.3 Let S ⊂ Rn be an m-dimensional manifold, and let σ: U →
S be a chart. It follows from Theorem 1.7 that σ −1 is smooth σ(U ) → Rm .
Lemma 2.6. Let p ∈ X ⊂ Rn and Y ⊂ Rm . If f : X → Y is smooth at p,
and g: Y → Rl is smooth at f (p), then g ◦ f : X → Rl is smooth at p.
Proof. Let F : W → Rm be a local smooth extension of f around p. Likewise
let G: V → Rl be a local smooth extension of g around f (p) ∈ Y . The set
W = F −1 (V ) is an open neighborhood of p, and G◦(F |W ) is a local smooth
extension of g ◦ f at p.
The definition of smoothness that we have given for manifolds in Rn uses
the ambient space Rn . In order to prepare for the generalization to abstract
manifolds, we shall now give an alternative description.
Theorem 2.6. Let f : S → Rl be a map and let p ∈ S. If f is smooth at p,
then f ◦ σ: U → Rl is smooth at σ −1 (p) for each chart (σ, U ) on S around p.
Conversely, if f ◦ σ is smooth at σ −1 (p) for some chart σ around p, then f
is smooth at p.
Here the smoothness of f ◦ σ refers to the ordinary notion as explained
above Definition 2.6.1.
Abstract manifolds 25
Proof. The first statement is immediate from Lemma 2.6. For the converse,
assume f ◦ σ is smooth at σ −1 (p) and apply Lemma 2.6 and Example 2.6.3
to f |σ(U) = (f ◦ σ) ◦ σ −1 . It follows that f |σ(U) is smooth at p. Since σ(U )
is relatively open in S (by Corollary 1.7), a local smooth extension of f |σ(U)
is also a local smooth extension of f , hence f is smooth at p.
p f f (p)
S ⊂ Rn S̃ ⊂ Rl
σ σ̃
U ⊂ Rm Ũ ⊂ Rk
σ̃ −1 ◦ f ◦ σ
Note that the statement that p is interior is of importance, since the set
(2.2), on which σ̃ −1 ◦f ◦σ is defined, then contains an open set around σ −1 (p),
so that the usual notion of smoothness can be applied to this map at this
point.
Proof. Assume that f is smooth at p. It was remarked below Definition 2.6.1
that then it is continuous at p. Since σ̃(Ũ ) is open in S̃ by Corollary 1.7, it
follows that f maps a neighborhood of p in S into σ̃(Ũ ), hence p is interior
as stated.
It follows from Theorem 2.6 that f ◦ σ is smooth at q = σ −1 (p), and from
Lemma 2.6 and Example 2.6.3 that the composed map σ̃ −1 ◦ f ◦ σ is smooth
at q.
For the converse we just have to note that the identity
f ◦ σ = σ̃ ◦ (σ̃ −1 ◦ f ◦ σ),
σ̃ −1 ◦ f ◦ σ,
Abstract manifolds 27
which is defined on (2.2), the coordinate expression for f with respect to the
given charts.
Example 2.8.3 Let G = SO(2), the group of all 2 × 2 real matrices which
are orthogonal, that is, they satisfy the relation AAt = I, and which have
determinant 1. The set G is in one-to-one correspondence with the unit circle
in R2 by the map
x1 x2
(x1 , x2 ) → .
−x2 x1
If we give G the smooth structure so that this map is a diffeomorphism,
then it becomes a 1-dimensional Lie group, called the circle group. The
multiplication of matrices is given by a smooth expression in x1 and x2 , and
so is the inversion x → x−1 , which only amounts to a change of sign on x2 .
Example 2.8.4 Let G = GL(n, R), the set of all invertible n × n matrices.
It is a group, with matrix multiplication as the operation. It is a manifold
in the following fashion. The set M(n, R) of all real n × n matrices is in
2
bijective correspondence with Rn and is therefore a manifold of dimension
n2 . The subset G = {A ∈ M(n, R) | det A
= 0} is an open subset, because
the determinant function is continuous. Hence G is a manifold.
Furthermore, the matrix multiplication M(n, R) × M(n, R) → M(n, R) is
given by smooth expressions in the entries (involving products and sums),
Abstract manifolds 29
Rn there is a countable base for the topology. The same is then true for any
subset X ⊂ Rn , since the collection of intersections with X of elements from
a base for Rn , is a base for the topology in X.
Lemma 2.9. Let M be a differentiable manifold. Then M has a countable
atlas if and only if there exists a countable base for the topology.
A topological space, for which there exists a countable base, is said to be
second countable.
Proof. Assume that M has a countable atlas. For each chart σ: U → M in
the atlas there is a countable base for the topology of U , according to the
example above, and since σ is a homeomorphism it carries this to a countable
base for σ(U ). The collection of all these sets for all the charts in the atlas,
is then a countable base for the topology of M , since a countable union of
countable sets is again countable.
Assume conversely that there is a countable base (Vk )k∈I for the topology.
For each k ∈ I, we select a chart σ: U → M for which Vk ⊂ σ(U ), if such
a chart exists. The collection of selected charts is clearly countable. It
covers M , for if x ∈ M is arbitrary, there exists a chart σ (not necessarily
among the selected) around x, and there exists a member Vk in the base with
x ∈ Vk ⊂ σ(U ). This member Vk is contained in a chart, hence also in a
selected chart, and hence so is x. Hence the collection of selected charts is a
countable atlas.
Corollary 2.9. Let S be a manifold in Rn . There exists a countable atlas
for S.
Proof. According to Example 2.9 there is a countable base for the topol-
ogy.
In Example 2.3.4 we introduced a 0-dimensional smooth structure on an
arbitrary set X, with the discrete topology. Any basis for the topology must
contain all singleton sets in X. Hence if the set X is not countable, there
does not exist any countable atlas for this manifold.
The theorem could give one the impression that if we limit our interest
to manifolds with a countable atlas, then the notion of abstract manifolds
is superfluous. This is not so, because in many circumstances it would be
very inconvenient to be forced to perceive a particular smooth manifold as a
subset of some high-dimensional Rk . The abstract notion frees us from this
limitation.
1.11 Exercises
Observe that the tangent space is a ‘local’ object, in the sense that if
two parametrized manifolds σ: U → Rn and σ : U → Rn are equal on some
neighborhood of a point x0 ∈ U ∩ U , then Tx0 σ = Tx0 σ .
From Geometry 1 we recall the following result, which is easily generalized
to the present setting.
Theorem 3.1. The tangent space is invariant under reparametrization. In
other words, if φ: W → U is a diffeomorphism of open sets in Rm and τ =
σ ◦ φ, then Ty0 τ = Tφ(y0 ) σ for all y0 ∈ W .
Proof. The proof is essentially the same as in Geometry 1, Theorem 2.7.
By the chain rule, the matrix identity Dτ (y0 ) = Dσ(φ(y0 ))Dφ(y0 ) holds.
Equivalently,
dτy0 (v) = dσφ(y0 ) (Dφ(y0 )v) (3.2)
for all v ∈ Rm . Hence Ty0 τ ⊂ Tφ(y0 ) σ. The opposite inclusion follows by the
same argument with φ replaced by its inverse.
34 Chapter 3
Tp S = {v ∈ Rn | Df (p)v = 0}.
dσx0 : Rm → T (3.6)
Every vector in T occurs in this fashion as the tangent vector for some
parametrized curve γ on M through p.
For the case M = S, compare the definition in (3.9) with (3.4).
Proof. If τ is another chart around p and ν = τ −1 ◦ γ the corresponding
coordinate expression, then
ν = τ −1 ◦ γ = (τ −1 ◦ σ) ◦ (σ −1 ◦ γ) = (τ −1 ◦ σ) ◦ μ
It now follows from (3.7) that dτy0 (ν (t0 )) = dσx0 (μ (t0 )).
For the last statement we essentially just repeat from the proof of Theorem
3.2. For each v ∈ Rm let γv (t) = σ(x0 + tv) for t close to 0. This is a
parametrized curve on M through p. In fact, in this case μ(t) = x0 + tv,
hence μ (0) = v and γv (0) = dσx0 (v). It follows that v → γv (0) is exactly
our isomorphism dσx0 of Rm onto T .
It follows that the curves γ1 = δw and γ2 (t) = γv obey the relation (3.11).
Hence γ1 ∼p γ2 and (3.12) follows.
(iii) This was established in the course of (ii).
We have introduced a vector structure on Tp M , but if M = S where S ⊂
R then Tp S already has such a structure, inherited from the surrounding Rn .
n
However, it follows from Lemma 3.3 that the two structures are isomorphic.
Example 3.5 Let V be an m-dimensional real vector space regarded as an
abstract manifold as in Example 2.3.1. We will show that for each element
p ∈ V there exists a natural isomorphism L: V → Tp V . For each vector
r ∈ V we define L(r) to be the ∼p -class of the line p + tr with direction r,
regarded as a curve through p. In order to prove that L is an isomorphism of
vector spaces, we choose a basis v1 , . . . , vm for V , and let σ: Rm → V denote
the corresponding isomorphism (x1 , . . . , xm ) → x1 v1 + · · · + xm vm , which
constitutes a chart on V (see Example 2.3.1). For x ∈ Rm and r = σ(x) we
have, with the notation of Theorem 3.5,
and hence dσx0 (x) = [γx ]p = L(r). Since dσx0 is an isomorphism of vector
spaces Rm → Tp V (by Theorem 3.5(ii)), and since x → r is an isomorphism
Rm → V , we conclude that r → L(r) is an isomorphism V → Tp V .
d
Dp,X (f ) =
f (p + tX) ∈ R. (3.13)
dt t=0
40 Chapter 3
∂ ∂
DX f = [ + 3 ](xz)(p) = 1.
∂x ∂y
The desired identity of dfp with the expression (3.18) for τ and ζ now follows
by applying the chain rule (3.17) to
ζ −1 ◦ f ◦ τ = (ζ −1 ◦ η) ◦ (η −1 ◦ f ◦ σ) ◦ (σ −1 ◦ τ ).
The tangent space 43
Definition 3.8. The map dfp : Tp M → Tf (p) N in Lemma 3.8.1 is called the
differential of f .
The chain rule for manifolds reads as follows. If g: L → M and f : M → N
are smooth maps between differentiable manifolds, then
dfp (γ (t0 )) = dηu0 (d(η −1 ◦ f ◦ σ)x0 (μ (t0 ))) = dηu0 (λ (t0 )),
From this relation and the similar one with opposite order of f and f −1 , we
see that when f is a diffeomorphism, then dfp is bijective with (dfp )−1 =
d(f −1 )f (p) . We thus obtain that the differential of a diffeomorphism is a
linear isomorphism between tangent spaces. In particular, two manifolds
between which there exists a diffeomorphism, must have the same dimension.
for all ϕ ∈ C ∞ (N ).
σx 1 , σx 2
∂
Di f = (f ◦ σ)(x0 ). (3.21)
∂xi
∂
Because of (3.21) it is customary to denote Di by ∂xi
.
3.10 Orientation
Let V be a finite dimensional vector space. Two ordered bases (v1 , . . . , vn )
and (ṽ1 , . . . , ṽn ) are said to be equally oriented if the transition matrix S,
whose columns are the coordinates of the vectors v1 , . . . , vn with respect
to the basis (ṽ1 , . . . , ṽn ), has positive determinant. Being equally oriented
is an equivalence relation among bases, for which there are precisely two
equivalence classes. The space V is said to be oriented if a specific class has
been chosen, this class is then called the orientation of V , and its member
bases are called positive. The Euclidean spaces Rn are usually oriented by the
class containing the standard basis (e1 , . . . , en ). For the null space V = {0}
we introduce the convention that an orientation is a choice between the signs
+ and −.
Example 3.10.1 For a two-dimensional subspace V of R3 it is customary
to assign an orientation by choosing a normal vector N . The positive bases
(v1 , v2 ) for V are those for which (v1 , v2 , N ) is a positive basis for R3 (in
other words, it is a right-handed triple).
Let σ be a chart on a differentiable manifold M , then the tangent space
is equipped with the standard basis (see Section 3.8) with respect to σ. For
each p ∈ σ(U ) we say that the orientation of Tp M , for which the standard
basis is positive, is the orientation induced by σ.
46 Chapter 3
3.11 Exercises
1 Let
M = {(x, y, z) ∈ R3 | x2 + y 2 + z 2 = 1 and x2 = yz 2 }
and put C = M \ {(0, ±1, 0)} (see also Exercise 1.16).
a. Determine Tp C where p = (0, 0, 1).
b. Prove that C ∪ {(0, 1, 0)} is not a curve in R3 .
The tangent space 47
Submanifolds
4.1 Submanifolds in Rk
j(x1 , . . . , xn ) = (x1 , . . . , xn , 0, . . . , 0) ∈ Rm .
Rm
U σ
N σ(U )
M
R n p
of σ̃, and by shrinking the sets U and V . The modification will be of the
form σ = σ̃ ◦ Φ, with Φ a suitably defined diffeomorphism of a neighborhood
of 0 in Rm .
Let
Ψ = σ̃ −1 ◦ τ : Rn → Rm ,
then Ψ is defined in a neighborhood of 0, and we have Ψ(0) = 0. Notice that
we can view Ψ as the coordinate expression σ̃ −1 ◦ i ◦ τ for the inclusion map
in these coordinates around p.
The Jacobian matrix DΨ(0) of Ψ at 0 is an m × n matrix, and since dip
is injective this matrix has rank n. By a reordering of the coordinates in Rm
we may assume that
A
DΨ(0) =
B
where A is an n × n matrix with non-zero determinant, and B is an arbitrary
(m − n) × n matrix.
Define Φ: Rm → Rm on a neighborhood of 0 by Φ(x, y) = Ψ(x) + (0, y) for
x ∈ Rn , y ∈ Rm−n . The Jacobian matrix of Φ at 0 has the form
A 0
DΦ(0) =
B I
y 0
x
−1 critical
The height function on the sphere, see Example 4.4
already from the fact that these operations are continuous. Next we choose
a neighborhood W of e in G according to Corollary 4.3.1. It follows from the
continuity of the map π: W → H in that corollary, that the equality π(h) = h
holds also for all h ∈ W ∩ H̄. Hence W ∩ H̄ ⊂ H. Let g ∈ H̄ be given. Then
since multiplication is continuous, {y ∈ G | g −1 y ∈ W } is a neighborhood of
g and hence contains an element h ∈ H. Then g −1 h ∈ W ∩ H̄ ⊂ H, and we
conclude that g = h(g −1 h)−1 ∈ H.
By a remarkable theorem of Cartan, every closed subgroup of a Lie group
is an (embedded) Lie subgroup. The proof is too complicated to be given
here.
Example 4.5 If G = GL(n, R) then the first statement in Theorem 4.5
follows from Theorem 2.8.
AAt = I.
∂D
y
x
Submanifolds 59
D = Hm = {x ∈ Rm | xm > 0}
D = {x ∈ Rm | x2 < 1}
W ∩ ∂D = {x ∈ W | f (x) = 0}.
and the map x → σ(x, 0), defined for all x = (x1 , . . . , xm−1 ) ∈ Rm−1 with
(x, 0) = (x1 , . . . , xm−1 , 0) ∈ U ◦ , is a chart on ∂D.
60 Chapter 4
M
U+ σ
D
σ(U + ) = D ∩ σ(U )
◦
U
U−
4.10 Exercises
1 Verify the statement below Example 4.2.4, about transitivity of the prop-
erty of being a submanifold.
2 Let M, N be differentiable manifolds and fix an element y0 ∈ N . Prove
that the subset {(x, y0 ) | x ∈ M } is a submanifold of M × N .
3 Let M be a differentiable manifold. Prove that the diagonal {(x, x) | x ∈
M } is a submanifold of M × M .
4 Let f : M → N be a diffeomorphism of manifolds, and let A ⊂ M be a
submanifold. Show that f (A) is a submanifold in N .
Submanifolds 63
5.1 Compactness
Recall that in a metric space X, a subset K is said to be compact, if
every sequence from K has a subsequence which converges to a point in
K. Recall also that every compact set is closed and bounded, and that the
converse statement is valid for X = Rn with the standard metric, that is,
the compact subsets of Rn are precisely the closed and bounded ones.
The generalization of compactness to an arbitrary topological space X
does not invoke sequences. It originates from another important property
of compact sets in a metric space, called the Heine-Borel property, which
concerns coverings of K.
Let X be a Hausdorff topological space, and let K ⊂ X.
Definition 5.1.1. A covering of K is a collection of sets Ui ⊂ X, where
i ∈ I, whose union ∪i Ui contains K. A subcovering is a subcollection (Uj )j∈J ,
where J ⊂ I, which is again a covering of K. An open covering is a covering
by open sets Ui ⊂ X.
K1 ⊂ D1 ∪ D2 ∪ · · · ∪ Di1
K2 ⊂ D1 ∪ D2 ∪ · · · ∪ Di2
x
r s
ψ(s − t)
h(t) =
ψ(s − t) + ψ(t − r)
is smooth, and it takes the value 1 for t ≤ r and 0 for t ≥ s. The function
ϕ(x) = h(|x|) has the required property.
y y
x x
y = ψ(x) y = h(x)
We first show that F is injective. Let p ∈ M and choose i such that fi (p) > 0.
Then p ∈ supp fi ⊂ σi (Ui ) and hi (p) = fi (p)x with x = σ −1 (p) ∈ Ui . Hence
xi
Wi = {(x1 , . . . , xn , y1 , . . . , yn ) ∈ RN | yi > 0, ∈ Ui },
yi
5.7 Connectedness
In this section two different notions of connectedness for subsets of a topo-
logical space are introduced and discussed. Let X be a non-empty topological
space.
Definition 5.7. (1) X is said to be connected if it cannot be separated in
two disjoint non-empty open subsets, that is, if X = A1 ∪ A2 with A1 , A2
open and disjoint, then A1 or A2 is empty (and A2 or A1 equals X).
(2) X is called pathwise connected if for each pair of points a, b ∈ X there
exists real numbers α ≤ β and a continuous map γ: [α, β] → X such that
γ(α) = a and γ(β) = b (in which case we say that a and b can be joined by
a continuous path in X).
(3) A non-empty subset E ⊂ X is called connected or pathwise connected
if it has this property as a topological space with the induced topology.
The above definition of ”connected” is standard in the theory of topolog-
ical spaces. However, the notion of ”pathwise connected” is unfortunately
sometimes also referred to as ”connected”. The precise relation between the
two notions will be explained in this section and the following. The empty
set was excluded in the definition, let us agree to call it both connected and
pathwise connected.
Example 5.7.1 A singleton E = {x} ⊂ X is clearly both connected and
pathwise connected.
Example 5.7.2 A convex subset E ⊂ Rn is pathwise connected, since by
definition any two points from E can be joined by a straight line, hence a
continuous curve, inside E. It follows from Theorem 5.7.3 below that such a
subset is also connected.
Example 5.7.3 It is a well-known fact, called the intermediate value prop-
erty, that a continuous real function carries intervals to intervals. It follows
from this fact that a subset E ⊂ R is pathwise connected if and only if it is
an interval. We shall see below in Theorem 5.7.1 that likewise E is connected
if and only if it is an interval. Thus for subsets of R the two definitions agree.
Topological properties of manifolds 73
Ct
y
x
H0
interval [α, β] is connected, it follows from Theorem 5.7.2 that C = γ([α, β])
is connected. We have reached a contradiction.
The converse statement is false. There exists subsets of, for example Rn
(n ≥ 2), which are connected but not pathwise connected (an example in
R2 is given below). However, for open subsets of Rn the two notions of
connectedness agree. This will be proved in the following section.
Example 5.7.7 The graph of the function
sin(1/x) x
= 0
f (x) =
0 x=0
is connected but not pathwise connected.
5.9 Components
Let X be topological space. We shall determine a decomposition of X as
a disjoint union of connected subsets. For example, the set R× = R \ {0} is
the disjoint union of the connected subsets ] − ∞, 0[ and ]0, ∞[.
Definition 5.9. A component (or connected component) of X is a subset
E ⊂ X which is maximal connected, that is, it is connected and not properly
contained in any other connected subset.
Example 5.9.1 The components of Q (with the standard metric from R)
are the point sets {q}. This is easily seen.
In the example mentioned before Definition 5.9, the components ] − ∞, 0[
and ]0, ∞[ are both open and closed in R× . In general, components need
not be open (see Example 5.9.1), but they are always closed. This is a
consequence of the following lemma.
Lemma 5.9. The closure of a connected subset E ⊂ X is connected.
Proof. Assume that Ē is separated in a disjoint union Ē = A1 ∪ A2 where
A1 , A2 are relatively open in Ē. Then Ai = Wi ∩ Ē for i = 1, 2, where W1
and W2 are open in X. Hence each set Wi ∩ E = Ai ∩ E is relatively open in
E, and these two sets separate E in a disjoint union. Since E is connected,
one of the two sets is empty. If for example W1 ∩ E = ∅, then E is contained
in the complement of W1 , which is closed in X. Hence also Ē is contained
in this complement, and we conclude that A1 = Ē ∩ W1 is empty.
Theorem 5.9. X is the disjoint union of its components. If X is locally
pathwise connected, for example if it is a differentiable manifold, then the
components are open and pathwise connected.
Proof. Components are disjoint, because if two components overlapped, their
union would be connected by Lemma 5.7, and none of them would be max-
imal. Let x ∈ X be arbitrary. It follows from Lemma 5.7 that the union
of all the connected sets in X that contain x, is connected. Clearly this
union is maximal connected, hence a component. Hence X is the union of
its components.
Assume that X is locally pathwise connected, and let E ⊂ X be a compo-
nent of X. For each x ∈ E there exists a pathwise connected open neighbor-
hood of x in X. This neighborhood must be contained in E (by maximality
of E). Hence E is open. Now Lemma 5.8 implies that E is pathwise con-
nected.
x y
H−
for θ ∈ R. Then O(2) = {kθ } ∪ {lθ }, and these are the connected compnents.
The special orthogonal group SO(2) = {kθ } is the identity component.
The preceding example can be generalized to O(n) for all n ≥ 1.
Theorem 5.10. The special orthogonal group SO(n) is connected. The com-
ponents of O(n) are SO(n) and {g ∈ O(n) | det g = −1}.
Proof. The proof is by induction on n. The cases n = 1 and n = 2 are easy.
Before we begin the induction, we need some notation.
Let G = SO(n) and H = SO(n − 1). We consider the embedding of H in
G as the subgroup of (n − 1) × (n − 1)-blocks in the lower right corner:
1 0
K = {kh := ∈ G | h ∈ H}.
0 h
Fix an arbitrary element g0 ∈ O(n) with determinant −1. Then the set
{g ∈ O(n) | det g = −1} is the image of SO(n) by g → g0 g, and hence it
is connected. Hence O(n) is the disjoint union of two connected subsets. If
these were not components of O(n), then O(n) would be connected. But
this is impossible, since {±1} is the image of O(n) by the determinant map,
which is continuous.
5.12 Exercises
m
Y (σ(u)) = ai (u)σu i (u) (6.1)
i=1
where σ(u) = p, and where ei ∈ Rm are the canonical basis vectors. These
are the analogs of the vectors σu i (u) in (6.1).
Definition 6.1.2. A vector field on M is an assignment of a tangent vector
Y (p) ∈ Tp M to each p ∈ M . It is called a smooth vector field if, for each
chart σ: U → M in a given atlas of M , there exist a1 , . . . , am in C ∞ (U ) such
that
m
Y (σ(u)) = ai (u)dσu (ei ) (6.3)
i=1
of the vector (dσu )−1 (Y (p)) ∈ Rm , where p = σ(u). Hence the condition of
smoothness can also be phrased as the condition that
is smooth U → Rm .
Example 6.1 Let M ⊂ Rm be an open subset, regarded as a manifold with
the chart of the identity map Rm → Rm . We know that at each point p ∈ M
the tangent space Tp M can be identified with Rm , and a smooth vector field
on M is nothing but a smooth map Y : M → Rm .
If we regard tangent vectors to Rm as directional differentiation operators
(see Section 3.6), the standard basis vectors are the partial derivative oper-
∂
ators ∂x i
, for i = 1, . . . , m. The operator corresponding to Y is the partial
differential operator
m
∂
ai (x) ,
i=1
∂xi
ξi (σ(u)) = ui , u ∈ U.
Lemma 6.2.1. The coordinate functions ξi belong to C ∞ (σ(U )). For each
X ∈ Tp M the coordinates of X with respect to the standard basis (6.2) are
Dp,X (ξ1 ), . . . , Dp,X (ξm ).
84 Chapter 6
∂
Dp,dσu (ei ) f = (f ◦ σ)(u) (6.5)
∂ui
(see (3.21)). If X = i ai dσu (ei ) ∈ Tp M we derive
∂ ∂
Dp,X (ξj ) = ai (ξj ◦ σ)(u) = ai (uj ) = aj .
i
∂ui i
∂ui
Notice that it follows from the preceding lemma, that a tangent vector
X ∈ Tp M is uniquely determined by its action on functions. If we know
Dp,X f for all smooth functions f defined on arbitrary neighborhoods of p,
then we can determine X by taking f = ξ1 , . . . , f = ξm with respect to some
chart. In particular, it follows that a vector field is uniquely determined by
its action on functions. Because of this, it is quite customary to identify a
smooth vector field Y on M with its action on smooth functions, and thus
regard the operator f → Y f as being the vector field itself. The following
lemma supports this point of view.
Lemma 6.2.2. Let Y be a vector field on an abstract manifold M . The
following conditions are equivalent:
(i) Y is smooth,
(ii) Y f ∈ C ∞ (M ) for all f ∈ C ∞ (M ),
(iii) Y f ∈ C ∞ (Ω) for all open sets Ω ⊂ M and all f ∈ C ∞ (Ω).
Proof. Let σ be a chart on M , and let (6.3) be the associated expression of
Y . It follows from (6.5) that
m
m
∂
Y f (σ(u)) = ai (u)Dσ(u),dσu (ei ) f = ai (u) (f ◦ σ)(u). (6.6)
i=1 i=1
∂ui
Y (f + g) = Y f + Y g, Y (λf ) = λY f
6.3 Derivations
Smooth vector fields obey a generalized Leibniz rule.
Definition 6.3. A linear map Y: C ∞ (M ) → C ∞ (M ) satisfying
The lemma now follows by applying the usual Leibniz rule to the differenti-
ation of the product [(f g) ◦ σ] = (f ◦ σ)(g ◦ σ).
In fact, the following converse is valid. The proof will not be given here.
Theorem 6.3. Every derivation Y of C ∞ (M ) is given by a unique vector
field Y ∈ X(M ).
for f ∈ C ∞ (Ω).
given by [X, Y ]f = XY f − Y Xf = dx d
f.
None of the operators XY and Y X are vector fields – they are second
order operators, whereas vector fields are first order operators. However, it
turns out that their difference [X, Y ] is again a vector field, the essential
reason being that the second order terms cancel with each other. This is
established in the following theorem.
Proof. Let p ∈ M . We will show that there exists a tangent vector Z(p) ∈
Tp M such that Z(p)f = ([X, Y ]f )(p) for all f ∈ C ∞ (Ω) where p ∈ Ω (as
remarked below Lemma 6.2.1 such a tangent vector, if it exists, is unique).
Choose a chart σ: U → M around p, and let
m
m
X(σ(u)) = ai (u)dσu (ei ), Y (σ(u)) = bj (u)dσu (ej )
i=1 j=1
m
∂ ∂ ∂ ∂
([X, Y ]f ) ◦ σ = ai bj (f ◦ σ) − bj ai (f ◦ σ)
i,j=1
∂ui ∂uj ∂uj ∂ui
m
∂bj ∂ ∂ai ∂
= ai (f ◦ σ) − bj (f ◦ σ).
∂ui ∂uj ∂uj ∂ui
i,j=1
Vector fields and Lie algebras 89
In the last step we used the Leibniz rule and cancellation. Let
m
∂bj ∂ai
Z(p) = ai dσu (ej ) − bj dσu (ei )
∂ui ∂uj
i,j=1
∂bi ∂ai
= aj − bj dσu (ei ) ∈ Tp M,
i j
∂uj ∂uj
The following lemma essentially shows that Lie brackets transform nicely
by smooth maps. Let ψ: M → N be a smooth map between differentiable
manifolds, and let X ∈ X(M ) and V ∈ X(N ). We shall say that dψ trans-
forms X into V if
dψp (X(p)) = V (ψ(p))
for all p ∈ M . Since vector fields are uniquely determined by their action on
smooth functions, an equivalent condition is obtained from Lemma 3.8.3,
X(ϕ ◦ ψ) = (V ϕ) ◦ ψ
for all ϕ ∈ C ∞ (N ).
90 Chapter 6
Proof. For (1), see Exercise 5. For (2) we apply Lemma 6.6.3. It is easily
seen that the inclusion map transforms X ∈ X(M ) into V ∈ X(N ) if and
only if X = V |M .
Example 6.7.2 Again, let G = Rn with addition. The Lie bracket of two
vector fields with constant coefficients,
∂ ∂
[X, Y ] = [ ai , bj ]=0
i
∂xi j ∂xj
[aX + bY, Z] = a[X, Z] + b[Y, Z], [X, aY + bZ] = a[X, Y ] + b[X, Z],
[X, Y ] = −[Y, X],
[[X, Y ], Z] + [[Y, Z], X] + [[Z, X], Y ] = 0,
fields, is not canonical. It can be shown that the use of right translation
instead of left translation would lead to an isomorphic Lie algebra (but in
general, it would be realized by a different set of vector fields).
Example 6.7.5 Let G = Rn with addition. It follows from Example 6.7.2
that g = Rn , and that the Lie bracket is trivial [X, Y ] = 0 for all X, Y ∈ g.
In the preceding example the Lie group G is commutative, and all the Lie
brackets [X, Y ] are trivial. This is not a coincidence, in fact it can be shown
that a connected Lie group is commutative if and only if its Lie algebra has
trivial brackets (we shall not prove this here). This is a simple instance of
the fact mentioned above, that a lot of information about the Lie group can
be retrieved from its Lie algebra.
The map F (x, y) = f (xy) is smooth on G × G, and the lemma shows that
g → X(e)(F (g, ·)) is smooth on G. This is exactly the desired conclusion.
Lemma 6.8.2. Let M, N be differentiable manifolds, and let F ∈ C ∞ (M ×
N ). For each x ∈ M , denote by F (x, ·) the smooth function y → F (x, y)
on N . Let q ∈ N and Y ∈ Tq N be fixed, then x → Y (F (x, ·)) is a smooth
function on M .
Proof. Let σ: U → M be a chart, and let γ: I → N be a parametrized curve
through q, which represents Y . The map (u, t) → F (σ(u), γ(t)) belongs to
C ∞ (U × I), hence so does its derivative with respect to t. As a function of
u, this derivative is Y (F (σ(u), ·)).
Theorem 6.8. Let G be a Lie group with Lie algebra g. The evaluation map
X → X(e) is a linear isomorphism of g onto the tangent space Te G.
Proof. The map is clearly linear, and it is injective, for if X(e) = 0 then
X(g) = (dg )e (X(e)) = 0 for all g ∈ G, hence X = 0. Finally, we show that
it is surjective. Let Y ∈ Te G be given, and define X(x) = d(x )e (Y ) for all
x ∈ G. Then X is a left invariant vector field on G, since by the chain rule
d
Xf (g) = f (g + tgA)|t=0 (6.8)
dt
when A = X(e).
We now turn to the actual proof. Let X, Y be left invariant vector fields,
and let A = X(e) and B = Y (e). Let T denote the left invariant vector field
for which T (e) = AB − BA. The claim is then that [X, Y ] = T , that is, we
need to show
[X, Y ]f = T f (6.9)
d
Xξ(g) = (ξ(g) + tξ(gA))|t=0 = ξ(gA). (6.10)
dt
d d
XY ξ(g) = Y ξ(g + tgA)|t=0 = ξ((g + tgA)B)|t=0 = ξ(gAB).
dt dt
Likewise Y Xξ(g) = ξ(gBA), and thus [X, Y ]ξ(g) = ξ(g(AB − BA)). Now
(6.9) follows.
Since O(n) is a Lie subgroup of GL(n, R), it follows from Corollary 6.9 that
o(n) is a Lie subalgebra of the Lie algebra of GL(n, R), and then it follows
from Theorem 6.10 that the Lie bracket of o(n) is the commutator bracket
[A, B] = AB − BA.
6.11 Exercises
1 Let Y denote the vector field Y (x, y) = (x, −y) on R2 . Determine its
coordinates with respect to the chart σ(u, v) = (u + v, u − v).
2 Prove that Y (x, y, z) = (xz, yz, z 2 − 1) is a smooth vector field on the
sphere S 2 .
3 Consider M = R2 . Compute the Lie bracket [X, Y ] when
a) X(x, y) = (x, 0), Y (x, y) = (0, y)
b) X(x, y) = (0, x), Y (x, y) = (y, 0).
4 Consider M = R. Determine the Lie bracket of two arbitrary smooth
vector fields X(x) = f (x) and Y (x) = g(x).
5 Let N ⊂ M be a submanifold of a differentiable manifold, and let V ∈
X(M ). Assume that V (p) ∈ Tp N for all p ∈ N . Prove that the restriction
V |N is a smooth vector field on N .
96 Chapter 6
AB = {xy | x ∈ A, y ∈ B}.
Dn = W n = {x1 · · · xn | x1 , . . . , xn ∈ W }.
Tensors
ξi (a1 e1 + · · · + an en ) = ai , a1 , . . . , an ∈ R.
(ii) The elements ξ1 , . . . , ξn form a basis for V ∗ (called the dual basis).
Proof. (i) is easy. For (ii), notice first that two linear forms on a vector space
are equal, if they agree on each element of a basis. Notice also that it follows
from the definition of ξi that ξi (ej ) = δij . Let ξ ∈ V ∗ , then
n
ξ= ξ(ei )ξi , (7.1)
i=1
Proof. For simplicity we assume dim V < ∞ (although the result is true in
general). Assume v
= 0. Then there exists a basis e1 , . . . , en for V with
e1 = v. The element ξ1 of the dual basis satisfies ξ1 (v) = 1.
The dual space is useful for example in the study of subspaces of V . This
can be seen from the following theorem, which shows that the elements of V ∗
can be used to detect whether a given vector belongs to a given subspace.
Proof. The proof is similar to the previous one, which corresponds to the
special case U = {0}. As before we assume dim V < ∞. Let e1 , . . . , em be
a basis for U , let em+1 = v, and extend to a basis e1 , . . . , en for V . The
element ξm+1 of the dual basis satisfies ξm+1 |U = 0 and ξm+1 (v) = 1.
The term ‘dual’ suggests some kind of symmetry between V and V ∗ . The
following theorem indicates such a symmetry for the finite dimensional case.
We define V ∗∗ = (V ∗ )∗ as the dual of the dual space.
Φ(v)(ξ) = ξ(v)
Proof. It is easily seen that Φ maps into V ∗∗ , and that it is linear. If Φ(v) = 0
then ξ(v) = 0 for all ξ, hence v = 0 by Lemma 7.1.1. Thus Φ is injective.
If dim V < ∞ then dim V ∗∗ = dim V ∗ = dim V , and hence Φ is also surjec-
tive.
7.3 Tensors
We now proceed to define tensors. Let k ∈ N. Given a collection of
vector spaces V1 , . . . , Vk one can define a vector space V1 ⊗ · · · ⊗ Vk , called
their tensor product. The elements of this vector space are called tensors.
However, we do not need to work in this generality, and we shall be content
with the situation where the vector spaces V1 , . . . , Vk are all equal to the
same space. In fact, the tensor space T k V we define below corresponds to
V ∗ ⊗ · · · ⊗ V ∗ in the general notation.
Definition 7.3.1. Let V k = V × · · · × V be the Cartesian product of k
copies of V . A map ϕ from V k to a vector space U is called multilinear if it
is linear in each variable separately (i.e. with the other variables held fixed).
is a k-tensor.
(R ⊗ S) ⊗ T = R ⊗ (S ⊗ T )
0 = ϕ(v1 + v2 , v1 + v2 , . . . )
= ϕ(v1 , v1 , . . . ) + ϕ(v1 , v2 , . . . ) + ϕ(v2 , v1 , . . . ) + ϕ(v2 , v2 , . . . )
= ϕ(v1 , v2 , . . . ) + ϕ(v2 , v1 , . . . ).
Alt(T ) = T, T ∈ A0 (V ) = R or T ∈ A1 (V ) = V ∗ .
and hence
1
Alt(η1 ⊗ · · · ⊗ ηk )(v1 , . . . , vk ) = det[(ηi (vj ))ij ]. (7.6)
k!
Let e1 , . . . , en be a basis for V , and ξ1 , . . . , ξn the dual basis for V ∗ . We
saw in Theorem 7.3 that the elements ξi1 ⊗ · · · ⊗ ξik form a basis for T k (V ).
We will now exhibit a similar basis for Ak (V ). We have seen already that
Ak (V ) = 0 if k > n.
Theorem 7.4. Assume k ≤ n. For each subset I ⊂ {1, . . . , n} with k
elements, let 1 ≤ i1 < · · · < ik ≤ n be its elements, and let
n!
dim Ak (V ) = .
k!(n − k)!
104 Chapter 7
Proof. It follows from the last statement in Lemma 7.4.2 that Alt: T k (V ) →
Ak (V ) is surjective. Applying Alt to all the basis elements ξi1 ⊗ · · · ⊗ ξik for
T k (V ), we therefore obtain a spanning set for Ak (V ). It follows from (7.6)
that Alt(ξi1 ⊗ · · · ⊗ ξik ) = 0 if there are repetitions among the i1 , . . . , ik .
Moreover, if we rearrange the order of the numbers i1 , . . . , ik , the element
Alt(ξi1 ⊗ · · · ⊗ ξik ) is unchanged, apart from a possible change of the sign.
Therefore Ak (V ) is spanned by the elements ξI in (7.7).
Consider an arbitrary linear combination T = I aI ξI with coefficients
aI ∈ R. Let J = (j1 , . . . , jk ) where 1 ≤ j1 < · · · < jk ≤ n, then it follows
from (7.6) that
1/k! if I = J
ξI (ej1 , . . . , ejk ) =
0 otherwise
1
η1 ∧ η2 = (η1 ⊗ η2 − η2 ⊗ η1 ).
2
Since the operator Alt is linear, the wedge product depends linearly on
the factors S and T . It is more cumbersome to verify the associative rule for
∧. In order to do this we need the following lemma.
Lemma 7.5.1. Let S ∈ T k (V ) and T ∈ T l (V ). Then
Alt(Alt(S) ⊗ T ) = Alt(S ⊗ T ).
T (vσ(k+1) , . . . , vσ(k+l) )
1
= (k+l)!k! sgn(σ ◦ τ ) S(vσ(τ (1)) , . . . , vσ(τ (k)) )
σ∈G τ ∈H
T (vσ(τ (k+1)) , . . . , vσ(τ (k+l)) ).
(R ∧ S) ∧ T = R ∧ (S ∧ T ) = Alt(R ⊗ S ⊗ T ).
(R ∧ S) ∧ T = Alt(Alt(R ⊗ S) ⊗ T ) = Alt(R ⊗ S ⊗ T )
106 Chapter 7
and
R ∧ (S ∧ T ) = Alt(R ⊗ Alt(S ⊗ T )) = Alt(R ⊗ S ⊗ T ).
T1 ∧ · · · ∧ Tr = Alt(T1 ⊗ · · · ⊗ Tr )
1
η1 ∧ · · · ∧ ηk (v1 , . . . , vk ) = det[(ηi (vj ))ij ] (7.9)
k!
ζ ∧ η = −η ∧ ζ. (7.10)
T ∧ S = (−1)kl S ∧ T (7.11)
Proof. The identity (7.10) follows immediately from the fact that η ∧ ζ =
1
2 (η ⊗ ζ − ζ ⊗ η).
Since Ak (V ) is spanned by elements of the type S = η1 ∧ · · · ∧ ηk , and
A (V ) by elements of the type T = ζ1 ∧ · · · ∧ ζl , where ηi , ζj ∈ V ∗ , it suffices
l
7.7 Exercises
S = ξ1 ⊗ ξ1 + ξ1 ⊗ ξ2 − ξ2 ⊗ ξ1 − ξ2 ⊗ ξ2
is pure, then ad − bc = 0.
c. Find a 2-tensor which is not pure. Is the set of pure 2-tensors a linear
subspace in T 2 (V )?
4 For two vector spaces V, W , we denote by Hom(V, W ) the space of lin-
ear maps from V to W . For example, V ∗ = Hom(V, R). Determine an
isomorphism between T k (V ) and Hom(V, T k−1 (V )).
108 Chapter 7
Suppose ξ˜1 . . . , ξ˜n is a second basis for V ∗ . Show that the bases ξ1 , . . . , ξn
and ξ˜1 . . . , ξ̃n are equally oriented if and only if ξ˜1 ∧ · · · ∧ ξ˜n = cξ1 ∧ · · ·∧ ξn
with c > 0.
Chapter 8
Differential forms
a covector
dfp ∈ Tp∗ M.
The change of notation is motivated by the desire to give dxi the precise
meaning, which it now obtains as the differential of xi at p. In order to
avoid the double subscript of d(xi )p , we denote this differential by dxi (p).
According to Example 8.1 it is a covector,
Thus, dxi (p) is the linear form on Tp M , which carries a vector to its i-th
coordinate in the basis (8.1).
Proof. The first statement is just a repetition of the fact that the standard
basis is a basis (see Section 3.9). For the second statement, we notice that
by the chain rule
∂ui
dxi (p)(dσu (ej )) = d(xi ◦ σ)u (ej ) = = δij , (8.3)
∂uj
since xi ◦ σ(u) = ui .
a field of covectors
ξ(p) ∈ Tp∗ M
m
∂(f ◦ σ)
df = dxi . (8.5)
i=1
∂ui
∂f ∂f
df = dx1 + dx2 ,
∂x1 ∂x2
∂a1 ∂a2
= . (8.7)
∂x2 ∂x1
Differential forms 113
ξi1 ∧ · · · ∧ ξik
ω(p) ∈ Ak (Tp M )
to each p ∈ M .
In particular, given a chart σ: U → M , the elements dxi1 ∧ · · · ∧ dxik ,
where 1 ≤ i1 < · · · < ik ≤ m, are k-forms on the open subset σ(U ) of M .
For each p ∈ σ(U ), a basis for Ak (Tp M ) is obtained from these elements.
Therefore, every k-form ω on M has a unique expression on σ(U ),
ω= aI dxi1 ∧ · · · ∧ dxik
I={i1 ,...,ik }
where aI : σ(U ) → R.
Definition 8.3.2. We call ω smooth if all the functions aI are smooth, for
each chart σ in an atlas of M . A smooth k-form is also called a differential
k-form. The space of differential k-forms on M is denoted Ak (M ).
In particular, since A0 (Tp M ) = R, we have A0 (M ) = C ∞ (M ), that is, a
differential 0-form on M is just a smooth function ϕ ∈ C ∞ (M ). Likewise, a
differential 1-form is nothing but a smooth covector field.
114 Chapter 8
f ∗ (ϕω) = (ϕ ◦ f )f ∗ ω, (8.9)
f ∗ (θ ∧ ω) = f ∗ (θ) ∧ f ∗ (ω), (8.10)
∗
f (dϕ) = d(ϕ ◦ f ). (8.11)
for v1 , . . . , vk ∈ Tp M ⊂ Tp N .
The following result will be important later.
Lemma 8.4.2. Let f : M → N be a smooth map between two manifolds of
the same dimension k. Let x1 , . . . , xk denote the coordinate functions of a
chart σ on M , and let y1 , . . . , yk denote the coordinate functions of a chart
τ on N . Then
where det(df ) in p is the determinant of the matrix for dfp with respect to
the standard bases for Tp M and Tf (p) N given by the charts.
Proof. The element f ∗ (dy1 ∧ · · · ∧ dyk ) is a k-form on M , hence a constant
multiple of dx1 ∧ · · · ∧ dxk (recall that dim Ak (V ) = 1 when dim V = k).
We determine the constant by applying f ∗ (dy1 ∧ · · · ∧ dyk ) to the vector
(dσp (e1 ), . . . , dσp (ek )) in (Tp M )k . We obtain, see (7.6),
On the other hand, applying dx1 ∧ · · · ∧ dxk to the same vector in (Tp M )k ,
we obtain, again by (7.6),
1 1
(dx1 ∧ · · · ∧ dxk )(dσp (e1 ), . . . , dσp (ek )) = det(δij ) = ,
k! k!
since dx1 , . . . , dxk is the dual basis. It follows that the constant we are
seeking is the determinant det(dyi (dfp (dσp (ej ))))
Notice that dyi (dfp (dσp (ej ))) are exactly the matrix elements for dfp . The
formula (8.12) now follows.
for ω ∈ Ak (M ) and for all σ. We need to verify that the expressions (8.14)
agree on overlaps of charts. Before we commence on this we shall prove that
dσ satisfies (a)-(f) above, with M replaced by σ(U ) everywhere.
(a) Let k = 0 and ω = f ∈ C ∞ (σ(U )). The only summand in the
expression (8.13) corresponds to I = ∅, and the coefficient a∅ is equal to f
itself. The expression (8.14) then reads dσ ω = df .
(b) The coefficients aI in (8.13) depend linearly on ω, and the expression
(8.14) is linear with respect to these coefficients, hence dσ ω depends linearly
on ω.
(c) Let ϕ ∈ C ∞ (σ(U )), ω ∈ Ak (σ(U )), and write ω according to (8.13). In
the corresponding expression for the product ϕω each coefficient aI is then
multiplied by ϕ, and hence definition (8.14) for this product reads
dσ (ϕω) = d(ϕaI ) ∧ dxi1 ∧ · · · ∧ dxik
I
= (aI dϕ + ϕ daI ) ∧ dxi1 ∧ · · · ∧ dxik
I
= dϕ ∧ aI dxi1 ∧ · · · ∧ dxik + ϕ daI ∧ dxi1 ∧ · · · ∧ dxik
I I
= dϕ ∧ ω + ϕ dσ ω
where the rule in Lemma 8.2.2 was applied in the second step.
(d) Notice first that if I = (i1 , . . . , ik ) with 1 ≤ i1 < · · · < ik ≤ m then by
definition
dσ (a dxi1 ∧ · · · ∧ dxik ) = da ∧ dxi1 ∧ · · · ∧ dxik (8.16)
for a ∈ C ∞ (σ(U )). The expression (8.16) holds also without the assumption
that the indices i1 , . . . , ik are increasingly ordered, for the dxi1 , . . . , dxik can
118 Chapter 8
be reordered at the cost of the same change of sign on both sides. The
expression also holds if the indices are not distinct, because in that case both
sides are 0.
Let ω1 ∈ Ak (σ(U )), ω2 ∈ Al (σ(U )), and write both of these forms as
follows, according to (8.13):
ω1 = aI dxi1 ∧ · · · ∧ dxik , ω2 = bJ dxj1 ∧ · · · ∧ dxjl ,
I J
= d ω1 ∧ ω2 + (−1)k ω1 ∧ dσ ω2
σ
where the sign (−1)k in the last step comes from passing dbJ to the right
past the elements dxi1 up to dxik .
(e) is proved first for k = 1. In this case f ∈ C ∞ (σ(U )) and
m
∂(f ◦ σ)
σ
d f = df = dxi .
i=1
∂xi
Hence
m
∂(f ◦ σ) m m
∂ 2 (f ◦ σ)
σ σ
d (d f ) = d ∧ dxi = dxj ∧ dxi .
i=1
∂x i i=1 j=1
∂x j ∂x i
2 2
The latter expression vanishes because the coefficient ∂∂x(fj ∂x◦σ)
i
= ∂∂x(fi ∂x
◦σ)
j
oc-
curs twice with opposite signs.
The general case is now obtained by induction on k. It follows from
property (d) that dσ (df1 ∧ · · · ∧ dfk ) can be written as a sum of two terms
Differential forms 119
by property (e).
Having established these properties for the operators dσ on σ(U ), we pro-
ceed with the proof that for two different charts σ, σ , the corresponding
operators dσ and dσ agree on the overlap σ(U ) ∩ σ (U ).
Assume first that U ⊂ U and σ = σ|U . In the expression (8.13) for
ω on σ (U ) we then have the restriction of each aI to this set, and since
d(aI |σ (U ) ) = daI |σ (U ) it follows that the expression (8.14) for dσ ω on
σ (U ) is just the restriction of the same expression for dσ ω. Hence dσ ω =
dσ ω on σ (U ) as claimed.
For the general case the observation we just made implies that we can
replace U by the open subset σ −1 (σ (U ) ∩ σ(U )) and σ by its restriction to
this subset, without affecting dσ ω on σ (U ) ∩ σ(U ). Likewise, we can replace
U by σ −1 (σ (U ) ∩ σ(U )). The result is that we may assume σ(U ) = σ (U ).
We now apply the rules (a)-(f) for dσ to the expression
ω= aI dxi1 ∧ · · · ∧ dxik
I
for ω with respect to the chart σ . It follows from (b), (c) and (e) that
dσ ω = dσ (aI dxi1 ∧ · · · ∧ dxik )
I
= daI ∧ dxi1 ∧ · · · ∧ dxik ,
I
and hence dσ ω = dσ ω as claimed.
Finally we can define the global operator d. Let p ∈ M be given and
choose a chart σ around p. We define dω(p) for each ω ∈ Ak (M ) by the
expression in Definition 8.5. It follows from what we just proved, that dω(p)
is independent of the choice of chart (in particular, it is also independent of
the atlas).
It is now clear that dω is a well defined smooth k + 1-form satisfying
(8.15). The properties (a)-(f) for d follow immediately from (8.15) and the
corresponding properties for dσ .
As in Section 7.6 we define
A(M ) = A0 (M ) ⊕ A1 (M ) ⊕ · · · ⊕ Am (M )
120 Chapter 8
where m = dim M . The elements of A(M ) are called differential forms. Thus
a differential form is a map which associates to each point p ∈ M a member
of the exterior algebra A(Tp M ), in a smooth manner. The operators d of
Theorem 8.5 can then be combined in a single linear operator, also denoted
d: A(M ) → A(M ).
g ∗ : Ak (N ) → Ak (M )
dM ◦ g ∗ = g ∗ ◦ dN ,
where dM and dN denotes the exterior differentiation for the two manifolds.
Proof. By Lemma 8.4.1, and properties (b), (c) and (e) above,
d(g ∗ ω) = d( (aI ◦ g) d(yi1 ◦ g) ∧ · · · ∧ d(yik ◦ g))
I
= d(aI ◦ g) ∧ d(yi1 ◦ g) ∧ · · · ∧ d(yik ◦ g).
I
On the other hand, by the properties of g ∗ listed in the proof of Lemma 8.4.1,
g ∗ (dω) = g ∗ ( daI ∧ dyi1 ∧ · · · ∧ dyik )
I
= d(aI ◦ g) ∧ d(yi1 ◦ g) ∧ · · · ∧ d(yik ◦ g).
I
8.8 Exercises
2xy dx + (x2 + y) dy
is closed on R2 \{(0, 0)}, and that it is exact on M but not on R2 \{(0, 0)}.
5 Let ξ be a covector field on an abstract manifold M . Prove that ξ is
smooth if ξ(Y ) ∈ C ∞ (M ) for all Y ∈ X(M ) (thus improving the ‘if’ of
Lemma 8.2.1).
6 Prove Lemma 8.3.
7 Let g: R2 → R3 be given by g(u, v) = (cos u cos v, cos u sin v, sin v). Deter-
mine g ∗ (xdz) and g ∗ (zdx ∧ dy).
8 Let M = R3 . Let ω1 = adx and ω2 = adx ∧ dy, where a ∈ C ∞ (M ).
Determine dω1 and dω2 . Compute the following differential forms
9 Verify Theorem 8.7 by direct computation in the case where g is the map
in exercise 7, and the operators are applied to the differential form yzdx.
124 Chapter 8
Chapter 9
Integration
The purpose of this chapter is to define integration on smooth manifolds,
and establish its relation with the differentiation operator d of the previous
chapter.
9.2 Integration on Rn
Recall from Geometry 1, Chapter 3, that a non-empty subset D ⊂ R2 is
called an elementary domain if it is compact and if its boundary is a finite
Integration 127
cf dA = c f dA
D
D
| f dA | ≤ |f | dA
D
D
f dA = f dA + f dA,
D1 ∪D2 D1 D2
where in the last line D1 and D2 are domains of integration with disjoint
interiors.
There is an important rule for change of variables, which reads as follows.
It will not be proved here.
128 Chapter 9
This integral is independent of the chart, for the same reason as before.
These considerations can be generalized to an m-dimensional smooth sur-
face in Rk . The factor (EG − F 2 )1/2 in the integral over D is generalized as
the square root of the determinant det(σu i · σu j ), i, j = 1, . . . , m, for a given
chart σ. The dot is the scalar product from Rn .
When trying to generalize further to an abstract m-dimensional manifold
M we encounter the problem that in general there is no dot product on Tp M ,
hence there is no way to generalize E, F and G. As a consequence, it is not
possible to define the integral of a function on M in a way that is invariant
under changes of of charts. One way out is to introduce the presence of an
inner product on tangent spaces as an extra axiom in the definition of the
concept of a manifold - this leads to so-called Riemannian geometry, which is
the proper abstract framework for the theory of the first fundamental form.
Another solution is to replace the integrand including (EG − F 2 )1/2 by a
differential form of the highest degree. This approach, which we shall follow
here, leads to a theory of integration for m-forms, but not for functions.
In particular, it does not lead to a definition of area, since that would be
obtained from the integration of the constant 1, which is a function, not a
form. However, we can still view the theory as a generalization to the theory
for surfaces, by defining the integral of a function f over a surface as the
integral of f times the volume form (see Example 8.3).
Integration 129
Proof. The open subset σ̃(Ũ ) ∩ g −1 (σ(U )) of σ̃(Ũ ) contains R̃. Hence we can
replace σ̃ by its restriction to the preimage in Ũ of this set with no effect
on the integral over R̃. That is, we may assume that σ̃(Ũ ) ⊂ g −1 (σ(U )), or
equivalently, g(σ̃(Ũ )) ⊂ σ(U ).
130 Chapter 9
where the integrals on the right are defined with respect to the charts σα , as
in Definition 9.3. The sum is finite, since only finitely many ϕα are non-zero
on R.
Theorem 9.4. The definition given above is independent of the choice of
the charts σα and the subsequent choice of a partition of unity.
and since ϕα ϕ̃α̃ ω is supported inside the intersection R ∩ σα (Dα ) ∩ σα̃ (Dα̃ ),
the latter expression equals
ϕα ϕ̃α̃ ω
R∩σα (Dα )∩σ̃α̃ (Dα̃ )
α̃∈Ã α∈A
The integral over R ∩ σα (Dα ) ∩ σ̃α̃ (Dα̃ ) has the same value for the two charts
by Theorem 9.3. Hence the last expression above is symmetric with respect
to the partitions indexed by A and Ã, and hence the original sum has the
same value if the partition is replaced by the other one.
It is easily seen that M ω depends linearly on ω. Moreover, in analogy
with Lemma 9.3:
Lemma 9.4. Let g: M̃ → M be an orientation preserving diffeomorphism,
and R̃ ⊂ M̃ a domain of integration. Then R = g(R̃) ⊂ M is a domain
of integration. Furthermore, let ω be a continuous m-form with pull back
ω̃ = g ∗ ω. Then
ω̃ = ω.
R̃ R
Notice that the requirement on the map gi is just that its restriction to the
interior of Di is a chart. If gi itself is a chart, σ = gi , then the present
function
φi is identical with the function φ ◦ σ in (9.2), and the formula for
Ri
ω by means of gi∗ ω is identical with the one in Definition 9.3.
For example, the unit sphere S 2 is covered in this fashion by a single map
g: D → S 2 of spherical coordinates
with D = [−π/2, π/2] × [−π, π], and thus we can compute the integral of a
2-form over S 2 by means of its pull-back by spherical coordinates, in spite of
the fact that this is only a chart on a part of the sphere (the point being, of
course, that the remaining part is a null set).
134 Chapter 9
the point being that the right hand side is independent of the choice of R.
Theorem 9.6. Stokes’ theorem. Let ω be a differential m − 1-form on
M , and assume that Ω̄ ∩ supp ω is compact. Then
dω = ω. (9.7)
Ω ∂Ω
We may now assume that ω is supported inside σ(D◦ ) for one of these
charts. For if the result has been established in this generality, we can apply
it to ϕα ω for each element ϕα in a partition of unity as in Definition 9.4.2.
Since the set R is compact, only finitely many ϕα ’s are non-zero on it, and
as both sides of (9.7) depend linearly on ω, the general result then follows.
The equation (9.7) clearly holds if σ(U ) ⊂ M \ Ω̄, since then both sides
are 0. This leaves the other two cases, σ(U ) ∩ Ω = σ(U + ) and σ(U ) ⊂ Ω, to
be checked.
We may also assume that
∂f
dω = (−1)j−1 dx1 ∧ · · · ∧ dxm ,
∂xj
where the sign appears because dxj has been moved past the 1-forms from
dx1 up to dxj−1 .
Assume first that j < m. We will prove that then
dω = ω = 0.
Ω ∂Ω
By definition
∂f ∂(f ◦ σ)
dω = (−1) j−1
dx1 ∧ · · · ∧ dxm = (−1)j−1 dV,
Ω Ω ∂xj ∂uj
where the last integration takes place over the set {x ∈ D | σ(x) ∈ Ω̄}, that
is, over D ∩ U + or D, in the two cases.
In the integral over the rectangle D ∩ U + or D, we can freely interchange
the order of integrations over u1 , . . . , um . Let us take the integral over uj
first (innermost), and let us denote its limits by a and b. Notice that since
j < m these are the limits for the uj variable both in D ∩ U + and D. Now
by the fundamental theorem of calculus
b
∂(f ◦ σ)
duj = f (σ(u1 , . . . , uj−1 , b, uj+1 , . . . , um ))
a ∂uj
− f (σ(u1 , . . . , uj−1 , a, uj+1 , . . . , um )).
However, since ω is supported in σ(D◦ ), it follows that these values are zero
for all u1 , . . . uj−1 , uj+1 , . . . um , and hence Ω dω = 0 as claimed.
136 Chapter 9
On the other hand, if σ(U )∩Ω = σ(U + ), then xm = 0 along the boundary,
and it follows immediately
that dxm , and hence also ω, restricts to zero on
∂Ω. Therefore also ∂Ω ω = 0. The same conclusion holds trivially in the
other case, σ(U ) ⊂ Ω.
Next we assume j = m. Again we obtain 0 on both sides of (9.7) in the
case σ(U ) ⊂ Ω, and we therefore assume the other case. We will prove that
then
m
dω = ω = (−1) f (σ(u1 , . . . , um−1 , 0)) du1 . . . dum−1 ,
Ω ∂Ω
where the integral runs over the set of (u1 , . . . , um−1 ) ∈ Rm−1 for which
σ(u1 , . . . , um−1 , 0) ∈ ∂Ω
Following the preceding computation of Ω dω, we take the integral over
uj = um first. This time, however, the lower limit a is replaced by the value
0 of xm on the boundary, and the previous conclusion fails, that f vanishes
here. Instead we obtain the value f (σ(u1 , . . . , um−1 , 0)), with a minus in
front because it is the lower limit in the integral. Recall that there was a
factor (−1)j−1 = (−1)m−1 in front. Performing the integral over the other
variables as well, we thus obtain the desired integral expression
m
(−1) f (σ(u1 , . . . , um−1 , 0)) du1 . . . dum−1
for dω.
Ω
On the other hand, the integral ∂Ω ω can be computed by means of the
restricted chart σ|U∩Rm−1 on ∂Ω. However, we have to keep track of the
orientation of this chart. By definition (see Section 4.8), the orientation of
the basis
dσ(e1 ), . . . , dσ(em−1 )
as claimed.
Integration 137
and
ω = f (b) − f (a)
∂Ω
Thus we see that in this case Stokes’ theorem reduces to the fundamental
theorem of calculus.
Example 9.7.2. Let M = R2 , then m = 2 and ω is a 1-form. The boundary
of the open set Ω is a union of smooth curves, we assume for simplicity it
is a single simple closed curve. Write ω = f (x, y) dx + g(x, y) dy, then (see
Example 8.6.2)
∂f ∂g
dω = (− + ) dx ∧ dy
∂y ∂x
and hence
∂f ∂g
dω = − + dA.
Ω Ω ∂y ∂x
On the other hand, the integral ∂Ω ω over the boundary can be computed
as follows. Assume γ: [0, T ] → ∂Ω is the boundary curve, with end points
γ(0) = γ(T ) (and no other self intersections). Theorem 9.5 can be applied
with D1 = [0, T ] and g1 = γ. Write γ(t) = (x(t), y(t)), then by definition
We thus see that in this case Stokes’ theorem reduces to the classical Green’s
theorem (which is equivalent with the divergence theorem for the plane):
∂f ∂g
− + dA = f (x, y)dx + g(x, y)dy.
Ω ∂y ∂x γ
138 Chapter 9
∂f ∂g ∂h
dω = ( + + ) dx ∧ dy ∧ dz
∂x ∂y ∂z
and hence
dω = div(f, g, h) dV,
Ω Ω
where div(f, g, h) = ∂f ∂g ∂h
∂x + ∂y + ∂z is the divergence of the vector field (f, g, h).
On the other hand, the integral ∂Ω ω over the boundary ∂Ω = S can be
computed as follows. SupposeD1 , . . .
, Dn and gi : Di → S are as in Theorem
n
9.5 for the manifold S. Then S ω = i=1 Di gi∗ ω. Let σ(u, v) = gi (u, v) be
one of the functions gi . Then
Furthermore
∂σ2 ∂σ2 ∂σ3 ∂σ3
d(y ◦ σ) ∧ d(z ◦ σ) = ( du + dv) ∧ ( du + dv)
∂u ∂v ∂u ∂v
∂σ2 ∂σ3 ∂σ2 ∂σ3
=( − ) du ∧ dv
∂u ∂v ∂v ∂u
and similarly
∂σ3 ∂σ1 ∂σ1 ∂σ3
d(z ◦ σ) ∧ d(x ◦ σ) = ( − ) du ∧ dv
∂u ∂v ∂v ∂u
∂σ1 ∂σ2 ∂σ2 ∂σ1
d(x ◦ σ) ∧ d(y ◦ σ) = ( − ) du ∧ dv.
∂u ∂v ∂v ∂u
Notice that the three expressions in front of du∧dv are exactly the coordinates
of the normal vector σu × σv . Thus we see that
Let N(u, v) denote the outward unit normal vector in σ(u, v), then σu ×σv =
σu × σv N and we see that D σ ∗ ω is the surface integral of the function
Integration 139
where
∂h ∂g ∂f ∂h ∂g ∂f
curl F = ( − , − , − ).
∂y ∂z ∂z ∂x ∂x ∂y
9.8 Exercises
2 Compute the integral of y dx + x dy, along the curve γ(t) = (sin7 t, sin17 t)
from t = 0 to t = π/2.
3 Let M = {(x, y, v, w)|x2 + y 2 = v 2 + w2 = 1} in R4 , and choose the
orientation as in Example 3.8.2. Let ω be the 2-form adx ∧ dv on R4 , with
a ∈ C ∞ (R4 ). Determine the integral of ω over M as an expression with
a, and compute it in each of the cases y 2 and a = yw.
4 On R2 , Stokes’ theorem is Green’s theorem (see Example 9.7.2).
a. Use the theorem to verify the following formulas for the area of Ω
1
A(Ω) = x dy = − y dx = x dy − y dx.
γ γ 2 γ
b. Show that
−y x 0 if(0, 0) ∈
/ Ω̄
2 2
dx + 2 dy =
γ x +y x + y2 2π if(0, 0) ∈ Ω
for all closed and regular smooth curves γ, where Ω is the interior (hint:
if (0, 0) ∈ Ω, use the theorem to replace γ by the unit circle).
5 On R3 , Stokes’ theorem is the divergence theorem (see Example 9.7.3)
a. Use (f, g, h) = (x, y, z) on the unit ball. What does the theorem tell?
b. Show that Green’s theorem can be reformulated as a 2-dimensional
version of the divergence theorem.
c. Find formulas for the volume of Ω, which are analogous to those of 4a
above.
d. Let
(x, y, z)
(f, g, h) = 2
(x + y 2 + z 2 )3/2
for which the divergence is zero. State and prove a result analogous the
that of 4b.
Index
algebraic variety, 11 diffeomorphism, 23, 27
alternating, 101 differentiable manifold, 20
form, 102 differential, 43,
ambient space, 10 form, 113
atlas, 8, 19 dimension of manifold, 1, 10, 20
compatible, 20 directional derivative, 39, 41, 40
countable, 29 discrete
base, for topology, 29 set, 10
basis, standard, 45 topology, 18
boundary, 18 divergence, div, 121
orientation, 60 theorem, 139
bump function, 68 domain
chain rule, 42, 43 of integration, 127, 131
chart, 8 with smooth boundary, 58
closed dual
set, 18 basis, 97
vector field, 112 map, 98
closure, 18 vector space, 97
compact, 65 Einstein convention, 101
compatible, 20 embedded
component, 76 parametrized manifold, 3
connected, 72 submanifold, 62
component, 76 embedding, 62
continuous, 18 exact vector field, 112
contravariant, 101 extension, local 23
coordinate exterior
expression, 34, 36 algebra, 107
function, 83, 110 differentiation, 116
map, 11 Gauss, 139
cotangent space, 109 GL(n, R), 28
countable, gl(n, R), 91, 94
atlas, 29 gradient, grad, 121
second, 30 Green, 137
covariant, 99 group, Lie, 28
covector, 109 orthogonal, 57
field , 111 Hausdorff topology, 19
cover, 8 Heine-Borel property, 65
covering, 65 homomorphism, Lie, 93
open, 65 homeomorphism, 3, 19
critical value, 55 immersed submanifold, 62
curl, 121 immersion, 62
curve, 5 induced topology, 19
142 Index
tangent
bundle, 86
map, 41
space, 34, 33, 38
tangential, 37
tensor, 99
contravariant, 101
covariant, 99
mixed, 101
product, 100
space, 99
topological space, 17
topology,
discrete, 18
Hausdorff, 19
trivial, 18
value,
critical, 55
regular, 55
variety, 11
vector,
cotangent, 109
field, 81
tangent, 38
volume form, 114
wedge product, 104
Whitney’s theorem, 30, 70