You are on page 1of 153

Manifolds and Differential Forms

Reyer Sjamaar
D EPARTMENT OF M ATHEMATICS , C ORNELL U NIVERSITY, I THACA , N EW Y ORK
14853-4201
E-mail address: 
  

URL:   !!!"#
$  
%
 &
'

Last updated: 2006-08-26T01:13-05:00


Copyright Reyer Sjamaar, 2001. Paper or electronic copies for personal use may
be made without explicit permission from the author. All other rights reserved.

Contents
Preface

Chapter 1. Introduction
1.1. Manifolds
1.2. Equations
1.3. Parametrizations
1.4. Configuration spaces
Exercises

1
1
7
9
9
13

Chapter 2. Differential forms on Euclidean space


2.1. Elementary properties
2.2. The exterior derivative
2.3. Closed and exact forms
2.4. The Hodge star operator
2.5. div, grad and curl
Exercises

17
17
20
22
23
24
27

Chapter 3. Pulling back forms


3.1. Determinants
3.2. Pulling back forms
Exercises

31
31
36
42

Chapter 4. Integration of 1-forms


4.1. Definition and elementary properties of the integral
4.2. Integration of exact 1-forms
4.3. The global angle function and the winding number
Exercises

47
47
49
51
53

Chapter 5. Integration and Stokes theorem


5.1. Integration of forms over chains
5.2. The boundary of a chain
5.3. Cycles and boundaries
5.4. Stokes theorem
Exercises

57
57
59
61
63
64

Chapter 6. Manifolds
6.1. The definition
6.2. The regular value theorem
Exercises

67
67
72
77

Chapter 7. Differential forms on manifolds

81

iii

iv

CONTENTS

7.1. First definition


7.2. Second definition
Exercises

81
82
89

Chapter 8. Volume forms


8.1. n-Dimensional volume in R N
8.2. Orientations
8.3. Volume forms
Exercises

91
91
94
96
100

Chapter 9. Integration and Stokes theorem on manifolds


9.1. Manifolds with boundary
9.2. Integration over orientable manifolds
9.3. Gau and Stokes
Exercises

103
103
106
108
109

Chapter 10. Applications to topology


10.1. Brouwers fixed point theorem
10.2. Homotopy
10.3. Closed and exact forms re-examined
Exercises

113
113
114
118
122

Appendix A. Sets and functions


A.1. Glossary
A.2. General topology of Euclidean space
Exercises

125
125
127
127

Appendix B. Calculus review


B.1. The fundamental theorem of calculus
B.2. Derivatives
B.3. The chain rule
B.4. The implicit function theorem
B.5. The substitution formula for integrals
Exercises

129
129
129
131
132
133
134

Bibliography

137

The Greek alphabet

139

Notation Index

141

Index

143

Preface
These are the lecture notes for Math 321, Manifolds and Differential Forms,
as taught at Cornell University since the Fall of 2001. The course covers manifolds and differential forms for an audience of undergraduates who have taken
a typical calculus sequence at a North American university, including basic linear algebra and multivariable calculus up to the integral theorems of Green, Gau
and Stokes. With a view to the fact that vector spaces are nowadays a standard
item on the undergraduate menu, the text is not restricted to curves and surfaces
in three-dimensional space, but treats manifolds of arbitrary dimension. Some
prerequisites are briefly reviewed within the text and in appendices. The selection of material is similar to that in Spivaks book [Spi65] and in Flanders book
[Fla89], but the treatment is at a more elementary and informal level appropriate
for sophomores and juniors.
A large portion of the text consists of problem sets placed at the end of each
chapter. The exercises range from easy substitution drills to fairly involved but,
I hope, interesting computations, as well as more theoretical or conceptual problems. More than once the text makes use of results obtained in the exercises.
Because of its transitional nature between calculus and analysis, a text of this
kind has to walk a thin line between mathematical informality and rigour. I have
tended to err on the side of caution by providing fairly detailed definitions and
proofs. In class, depending on the aptitudes and preferences of the audience and
also on the available time, one can skip over many of the details without too much
loss of continuity. At any rate, most of the exercises do not require a great deal of
formal logical skill and throughout I have tried to minimize the use of point-set
topology.
This revised version of the notes is still a bit rough at the edges. Plans for
improvement include: more and better graphics, an appendix on linear algebra, a
chapter on fluid mechanics and one on curvature, perhaps including the theorems
of Poincar-Hopf and Gau-Bonnet. These notes and eventual revisions can be
downloaded from the course website at
())*+-,,/...0#132
)(0-4567 899:08;<,=>/?213226,4
92 >>8>,@A3B, CD7 ;8
E$0F()1G9
.
Corrections, suggestions and comments will be received gratefully.
Ithaca, NY, 2006-08-26

CHAPTER 1

Introduction
We start with an informal, intuitive introduction to manifolds and how they
arise in mathematical nature. Most of this material will be examined more thoroughly in later chapters.
1.1. Manifolds
Recall that Euclidean n-space Rn is the set of all column vectors with n real
entries
IJ
LNM
J x1 M
J
M
x2
xH
. ,
K .. O
xn

which we shall call points or n-vectors and denote by lower case boldface letters. In
R2 or R3 we often write
I L
x
x
x HQP
,
resp.
x H K yO .
yR
z
For reasons having to do with matrix multiplication, column vectors are not to be
confused with row vectors S x 1 x2 . . . xn T . For clarity, we shall usually separate
the entries of a row vector by commas, as in S x 1 , x2 , . . . , xn T . Occasionally, to save
space, we shall represent a column vector x as the transpose of a row vector,
x HUS x1 , x2 , . . . , xn T T .
A manifold is a certain type of subset of Rn . A precise definition will follow
in Chapter 6, but one important consequence of the definition is that a manifold
has a well-defined tangent space at every point. This fact enables us to apply the
methods of calculus and linear algebra to the study of manifolds. The dimension of
a manifold is by definition the dimension of its tangent spaces. The dimension of
a manifold in Rn can be no higher than n.
Dimension 1. A one-dimensional manifold is, loosely speaking, a curve without kinks or self-intersections. Instead of the tangent space at a point one usually speaks of the tangent line. A curve in R2 is called a plane curve and a curve in
R3 is a space curve, but you can have curves in any Rn . Curves can be closed (as
in the first picture below), unbounded (as indicated by the arrows in the second
picture), or have one or two endpoints (the third picture shows a curve with an
endpoint, indicated by a black dot; the white dot at the other end indicates that
1

1. INTRODUCTION

that point does not belong to the curve; the curve peters out without coming to
an endpoint). Endpoints are also called boundary points.

A circle with one point deleted is also an example of a manifold. Think of a torn
elastic band.

By straightening out the elastic band we see that this manifold is really the same
as an open interval.
The four plane curves below are not manifolds. The teardrop has a kink, where
two distinct tangent lines occur instead of a single well-defined tangent line; the
five-fold loop has five points of self-intersection, at each of which there are two
distinct tangent lines. The bow tie and the five-pointed star have well-defined
tangent lines everywhere. Still they are not manifolds: the bow tie has a selfintersection and the cusps of the star have a jagged appearance which is proscribed
by the definition of a manifold (which we have not yet given). The points where
these curves fail to be manifolds are called singularities. The good points are
called smooth.

Singularities can sometimes be resolved. For instance, the self-intersections of


the Archimedean spiral, which is given in polar coordinates by r is a constant times

1.1. MANIFOLDS

, where r is allowed to be negative,

can be got rid of by uncoiling the spiral and wrapping it around a cone. You can
convince yourself that the resulting space curve has no singularities by peeking at
it along the direction of the x-axis or the y-axis. What you will see are the smooth
curves shown in the yz-plane and the xz-plane.

e3
e2
e1

(The three-dimensional models in these notes are drawn in central perspective.


They are best viewed facing the origin, which is usually in the middle of the picture, from a distance of 30 cm with one eye shut.) Singularities are extremely
interesting, but in this course we shall focus on gaining a thorough understanding
of the smooth points.

1. INTRODUCTION

Dimension 2. A two-dimensional manifold is a smooth surface without selfintersections. It may have a boundary, which is always a one-dimensional manifold. You can have two-dimensional manifolds in the plane R2 , but they are relatively boring. Examples are: an arbitrary open subset of R 2 , such as an open
square, or a closed subset with a smooth boundary.

A closed square is not a manifold, because the corners are not smooth.1

Two-dimensional manifolds in three-dimensional space include a sphere, a paraboloid and a torus.

e3
e2
e1

The famous Mbius band is made by pasting together the two ends of a rectangular
strip of paper giving one end a half twist. The boundary of the band consists of
1To be strictly accurate, the closed square is a topological manifold with boundary, but not a smooth
manifold with boundary. In these notes we will consider only smooth manifolds.

1.1. MANIFOLDS

two boundary edges of the rectangle tied together and is therefore a single closed
curve.

Out of the Mbius band we can create in two different ways a manifold without
boundary by closing it up along the boundary edge. According to the direction in
which we glue the edge to itself, we obtain the Klein bottle or the projective plane.
A simple way to represent these three surfaces is by the following diagrams. The
labels tell you which edges to glue together and the arrows tell you in which direction.
b
b
a

Mbius band

a
b
Klein bottle

b
projective plane

Perhaps the easiest way to make a Klein bottle is first to paste the top and bottom
edges of the square together, which gives a tube, and then to join the resulting
boundary circles, making sure the arrows match up. You will notice this cannot be
done without passing one end through the wall of the tube. The resulting surface
intersects itself along a circle and therefore is not a manifold.

A different model of the Klein bottle is found by folding over the edge of a Mbius
band until it touches the central circle. This creates a Mbius type band with a
figure eight cross-section. Equivalently, take a length of tube with a figure eight
cross-section and weld the ends together giving one end a half twist. Again the

1. INTRODUCTION

resulting surface has a self-intersection, namely the central circle of the original
Mbius band. The self-intersection locus as well as a few of the cross-sections are
shown in black in the following wire mesh model.

To represent the Klein bottle without self-intersections you need to embed it in


four-dimensional space. The projective plane has the same peculiarity, and it too
has self-intersecting models in three-dimensional space. Perhaps the easiest model
is constructed by merging the edges a and b shown in the gluing diagram for the
projective plane, which gives the following diagram.

a
a

a
a

First fold the lower right corner over to the upper left corner and seal the edges.
This creates a pouch like a cherry turnover with two seams labelled a which meet
at a corner. Now fuse the two seams to create a single seam labelled a. Below is a
wire mesh model of the resulting surface. It is obtained by welding together two
pieces along the dashed wires. The lower half shaped like a bowl corresponds to
the dashed circular disc in the middle of the square. The upper half corresponds
to the complement of the disc and is known as a cross-cap. The wire shown in
black corresponds to the edge a. The interior points of the black wire are ordinary
self-intersection points. Its two endpoints are qualitatively different singularities

1.2. EQUATIONS

known as pinch points, where the surface is crinkled up.

e3
e2

e1

1.2. Equations
Very commonly manifolds are given implicitly, namely as the solution set
of a system

1 V x1 , . . . , xn W$X

c1 ,

2 V x1 , . . . , xn W$X

c2 ,
..
.

m V x1 , . . . , xn W$X

cm ,

of m equations in n unknowns. Here 1 , 2 , . . . , m are functions, c 1 , c2 , . . . , cm


are constants and x 1 , x2 , . . . , xn are variables. By introducing the useful shorthand
YZ
\N]
YZ
\N]
YZ
\N]
Z x1 ]
Z 1 V x W ]
Z c1 ]
Z[
]
Z[
]
Z[
]
x2
2 V x W
c2
xX
V x W"X
,
cX
.. ,
.. ,
..
. ^
.^
. ^
xn

cn

m V x W

we can represent this system as a single equation

V x WX

c.

It is in general difficult to find explicit solutions of such a system. (On the positive
side, it is usually easy to decide whether any given point is a solution by plugging
it into the equations.) Manifolds defined by linear equations (i.e. where is a
matrix) are called affine subspaces of Rn and are studied in linear algebra. More
interesting manifolds arise from nonlinear equations.
1.1. E XAMPLE . Consider the system of two equations in three unknowns,
x2 _ y2 X 1,
y _ z X 0.
Here

V x W$Xa`

x2 _ y2
y_ z b

and

1
c Xc`
.
0b

1. INTRODUCTION

The solution set of this system is the intersection of a cylinder of radius 1 about
the z-axis (given by the first equation) and a plane cutting the x-axis at a 45 d angle
(given by the second equation). Hence the solution set is an ellipse. It is a manifold
of dimension 1.
1.2. E XAMPLE . The sphere of radius r about the origin in R n is the set of all x
in Rn satisfying the single equation e x egf r. Here

e x ehfji x k x fUl x21 m x22 m

k/k/k m x2n

is the norm or length of x and


xk yf

x1 y1 m x2 y2 m

k/k/k m xn yn

is the inner product or dot product of x and y. The sphere of radius r is an n n 1dimensional manifold in Rn . The sphere of radius 1 is called the unit sphere and
is denoted by Sn o 1 . What is a one-dimensional sphere? And a zero-dimensional
sphere?

The solution set of a system of equations may have singularities and is therefore not necessarily a manifold. A simple example is xy f 0, the union of the two
coordinate axes in the plane, which has a singularity at the origin. Other examples
of singularities can be found in Exercise 1.5.
Tangent spaces. Let us use the example of the sphere to introduce the notion
of a tangent space. Let
M fqp x r Rn s e x egf r t
be the sphere of radius r about the origin in Rn and let x be a point in M. There are
two reasonable, but inequivalent, views of how to define the tangent space to M
at
space at x consists of all vectors y such that
u x. The first view is that the tangent
y n x v:k x f 0, i.e. y k x f x k x f r 2 . In coordinates: y 1 x1 m k/k/k m yn xn f r2 . This
is an inhomogeneous linear equation in y. In this view, the tangent space at x is an
affine subspace of Rn , given by the single equation y k x f r 2 .
However, for most practical purposes it is easier to translate this affine subspace to the origin, which turns it into a linear subspace. This leads to the second
view of the tangent space at x, namely as the set of all y such that y k x f 0, and
this is the definition that we shall espouse. The standard notation for the tangent
space to M at x is Tx M. Thus
Tx M fjp y r Rn s y k x f

0t ,

a linear subspace of Rn . (In Exercise 1.6 you will be asked to find a basis of Tx M
for a particular x and you will see that Tx M is n n 1-dimensional.)
Inequalities. Manifolds with boundary are often presented as solution sets
of a system of equations together with one or more inequalities. For instance,
the closed ball of radius r about the origin in Rn is given by the single inequality
e x exw r. Its boundary is the sphere of radius r.

1.4. CONFIGURATION SPACES

1.3. Parametrizations
A dual method for describing manifolds is the explicit way, namely by parametrizations. For instance,
xy

cos ,

yy

sin

parametrizes the unit circle in R2 and


xy

cos cos ,

yy

sin cos ,

zy

sin

parametrizes the unit sphere in R3 . (Here is the angle between a vector and the
xy-plane and is the polar angle in the xy-plane.) The explicit method has various
merits and demerits, which are complementary to those of the implicit method.
One obvious advantage is that it is easy to find points lying on a parametrized
manifold simply by plugging in values for the parameters. A disadvantage is that
it can be hard to decide if any given point is on the manifold or not, because this
involves solving for the parameters. Parametrizations are often harder to come by
than a system of equations, but are at times more useful, for example when one
wants to integrate over the manifold. Also, it is usually impossible to parametrize a manifold in such a way that every point is covered exactly once. Such is
the case for the two-sphere. One commonly restricts the polar coordinates z , {
to the rectangle | 0, 2 }~| 2, 2} to avoid counting points twice. Only the
meridian y 0 is then hit twice, but this does not matter for many purposes, such
as computing the surface area or integrating a continuous function.
We will use parametrizations to give a formal definition of the notion of a
manifold in Chapter 6. Note however that not every parametrization describes a
manifold. Examples of parametrizations with singularities are given in Exercises
1.1 and 1.2.

1.4. Configuration spaces


Frequently manifolds arise in more abstract ways that may be hard to capture
in terms of equations or parametrizations. Examples are solution curves of differential equations (see e.g. Exercise 1.10) and configuration spaces. The configuration
of a mechanical system (such as a pendulum, a spinning top, the solar system, a
fluid, or a gas etc.) is its state or position at any given time. (The configuration
ignores any motions that the system may be undergoing. So a configuration is like
a snapshot or a movie still. When the system moves, its configuration changes.)
In practice one usually describes a configuration by specifying the coordinates of
suitably chosen parts of the system. The configuration space or state space of the system is an abstract space, the points of which are in one-to-one correspondence to
all physically possible configurations of the system. Very often the configuration
space turns out to be a manifold. Its dimension is called the number of degrees of
freedom of the system. The configuration space of even a fairly small system can be
quite complicated.

10

1. INTRODUCTION

1.3. E XAMPLE . A spherical pendulum is a weight or bob attached to a fixed


centre by a rigid rod, free to swing in any direction in three-space.

The state of the pendulum is entirely determined by the position of the bob. The
bob can move from any point at a fixed distance (equal to the length of the rod)
from the centre to any other. The configuration space is therefore a two-dimensional
sphere.
Some believe that only spaces of dimension 3 (or 4, for those who have
heard of relativity) can have a basis in physical reality. The following two examples show that this is not true.
1.4. E XAMPLE . Take a spherical pendulum of length r and attach a second one
of length s to the moving end of the first by a universal joint. The resulting system
is a double spherical pendulum. The state of this system can be specified by a pair of
vectors x, y , x being the vector pointing from the centre to the first weight and y
the vector pointing from the first to the second weight.

x
y

The vector x is constrained to a sphere of radius r about the centre and y to a sphere
of radius s about the head of x. Aside from this limitation, every pair of vectors
can occur (if we suppose the second rod is allowed to swing completely freely
and move through the first rod) and describes a distinct configuration. Thus
there are four degrees of freedom. The configuration space is a four-dimensional
manifold, known as the (Cartesian) product of two two-dimensional spheres.
1.5. E XAMPLE . What is the number of degrees of freedom of a rigid body moving in R3 ? Select any triple of points A, B, C in the solid that do not lie on one line.
The point A can move about freely and is determined by three coordinates, and
so it has three degrees of freedom. But the position of A alone does not determine
the position of the whole solid. If A is kept fixed, the point B can perform two

1.4. CONFIGURATION SPACES

11

independent swivelling motions. In other words, it moves on a sphere centred at


A, which gives two more degrees of freedom. If A and B are both kept fixed, the
point C can rotate about the axis AB, which gives one further degree of freedom.

C
A

The positions of A, B and C determine the position of the solid uniquely, so the
total number of degrees of freedom is 3 2 1 6. Thus the configuration space
of a rigid body is a six-dimensional manifold.
1.6. E XAMPLE (the space of quadrilaterals). Consider all quadrilaterals ABCD
in the plane with fixed sidelengths a, b, c, d.
D
c
C
d

b
A

(Think of four rigid rods attached by hinges.) What are all the possibilities? For
simplicity let us disregard translations by keeping the first edge AB fixed in one
place. Edges are allowed to cross each other, so the short edge BC can spin full
circle about the point B. During this motion the point D moves back and forth on
a circle of radius d centred at A. A few possible positions are shown here.

(As C moves all the way around, where does the point D reach its greatest leftor rightward displacement?) Arrangements such as this are commonly used in
engines for converting a circular motion to a pumping motion, or vice versa. The
position of the pump D is wholly determined by that of the wheel C. This
means that the configurations are in one-to-one correspondence with the points
on the circle of radius b about the point B, i.e. the configuration space is a circle.

12

1. INTRODUCTION

Actually, this is not completely accurate: for every choice of C, there are two
choices D and D for the fourth point! They are interchanged by reflection in the
diagonal AC.
D
C

D
So there is in fact another circles worth of possible configurations. It is not possible
to move continuously from the first set of configurations to the second; in fact they
are each others mirror images. Thus the configuration space is a disjoint union of
two circles.

This is an example of a disconnected manifold consisting of two connected components.


1.7. E XAMPLE (quadrilaterals, continued). Even this is not the full story: it is
possible to move from one circle to the other when b c a d (and also when
a b c d).
D
c
C
d

b
A

In this case, when BC points straight to the left, the quadrilateral collapses to a line
segment:

EXERCISES

13

and when C moves further down, there are two possible directions for D to go,
back up:

or further down:

This means that when b c


are merged at a point.

a d the two components of the configuration space

The juncture represents the collapsed quadrilateral. This configuration space is


not a manifold, but most configuration spaces occurring in nature are (and an
engineer designing an engine wouldnt want to use this quadrilateral to make a
piston drive a flywheel). More singularities appear in the case of a parallelogram
(a c and b d) and in the equilateral case (a b c d).

Exercises
1.1. The formulas x t sin t, y 1 cos t (t R) parametrize a plane curve. Graph
this curve as carefully as you can. You may use software and turn in computer output.
Also include a few tangent lines at judiciously chosen points. (E.g. find all tangent lines
with
slope 0, 1, and .) To compute tangent lines, recall that the tangent vector at a point

x, y of the curve has components dx dt and dy dt. In your plot, identify all points where
the curve is not a manifold.
1.2. Same questions as in Exercise 1.1 for the curve x

3at 1 t 3 , y

3at2 1 t3 .

1.3. Parametrize the space curve wrapped around the cone shown in Section 1.1.

14

1. INTRODUCTION

1.4. Sketch the surfaces defined by the following gluing diagrams.

a
b

a
c

b
d

a
a1

b2

a
a1

b2

b1

a2

b1

a2

a2

b1

a2

b1

b2

a1

b2

a1

(Proceed in stages, first gluing the as, then the bs, etc., and try to identify what you get
at every step. One of these surfaces cannot be embedded in R 3 , so use a self-intersection
where necessary.)
1.5. For the values of n indicated below graph the surface in R 3 defined by x n y2 z.
Determine all the points where the surface does not have a well-defined tangent plane.
(Computer output is OK, but bear in mind that few drawing programs do an adequate job
of plotting these surfaces, so you may be better off drawing them by hand. As a preliminary
step, determine the intersection of each surface with a general plane parallel to one of the
coordinate planes.)
(i)
(ii)
(iii)
(iv)

n
n
n
n

0.
1.
2.

3.

1.6. Let M be the sphere of radius n about the origin in Rn and let x be the point
1, 1, . . . , 1 on M. Find a basis of the tangent space to M at x. (Use that Tx M is the set of all
y such that y x 0. View this equation as a homogeneous linear equation in the entries
y1 , y2 , . . . , yn of y and find the general solution by means of linear algebra.)

1.7. What is the number of degrees of freedom of a bicycle? (Imagine that it moves
freely through empty space and is not constrained to the surface of the earth.)
1.8. Choose two distinct positive real numbers a and b. What is the configuration
space of all parallelograms ABCD such that AB and CD have length a and BC and AD
have length b? What happens if a b? (As in Examples 1.6 and 1.7 assume that the edge
AB is kept fixed in place so as to rule out translations.)
1.9. What is the configuration space of all pentagons ABCDE in the plane with fixed
sidelengths a, b, c, d, e? (As in the case of quadrilaterals, for certain choices of sidelengths
singularities may occur. You may ignore these cases. To reduce the number of degrees of

EXERCISES

15

freedom you may also assume the edge AB to be fixed in place.)

b
A

1.10. The Lotka-Volterra system is an early (ca. 1925) predator-prey model. It is the
pair of differential equations
dx
rx sxy,
dt
dy
py qxy,

dt
where x t represents the number of prey and y t the number of predators at time t, while
p, q, r, s are positive constants. In this problem we will consider the solution curves (also
called trajectories) x t , y t - of this system that are contained in the positive quadrant
(x 0, y 0) and derive an implicit equation satisfied by these solution curves. (The
Lotka-Volterra system is exceptional in this regard. Usually it is impossible to write down
an equation for the solution curves of a differential equation.)
(i) Show that the solutions of the system satisfy a single differential equation of the
form dy dx
f x g y , where f x is a function that depends only on x and

g y a function that depends only on y.


(ii) Solve the differential equation of part (i) by separating the variables, i.e. by writ1
ing
dy
f x dx and integrating both sides. (Dont forget the integration

g y
constant.)
(iii) Set p
q
r
s
1 and plot a number of solution curves. Indicate the

direction in which the solutions move. Be warned that solving the system may
give better results than solving the implicit equation! You may use computer
software such as Maple, Mathematica or MATLAB. A useful Java applet, 
'N ,
can be found at 'FND''FD'
N''''D .

CHAPTER 2

Differential forms on Euclidean space


The notion of a differential form encompasses such ideas as elements of surface area and volume elements, the work exerted by a force, the flow of a fluid, and
the curvature of a surface, space or hyperspace. An important operation on differential forms is exterior differentiation, which generalizes the operators div, grad
and curl of vector calculus. The study of differential forms, which was initiated by
E. Cartan in the years around 1900, is often termed the exterior differential calculus.
A mathematically rigorous study of differential forms requires the machinery of
multilinear algebra, which is examined in Chapter 7. Fortunately, it is entirely possible to acquire a solid working knowledge of differential forms without entering
into this formalism. That is the objective of this chapter.
2.1. Elementary properties
A differential form of degree k or a k-form on Rn is an expression

f I dx I .
I

(If you dont know the symbol , look up and memorize the Greek alphabet in
the back of the notes.) Here I stands for a multi-index i 1 , i2 , . . . , ik of degree k, that
is a vector consisting of k integer entries ranging between 1 and n. The f I are
smooth functions on Rn called the coefficients of , and dx I is an abbreviation for
dxi1 dxi2 / / dxik .
(The notation dxi1 dxi2 // dxik is also often used to distinguish this kind of
product from another kind, called the tensor product.)
For instance the expressions

sin x1 e x4 dx1 dx5 x2 x25 dx2 dx3 6 dx2 dx4 cos x2 dx5 dx3 ,
x1 x3 x5 dx1 dx6 dx3 dx2 ,

represent a 2-form on R5 , resp. a 4-form on R6 . The form consists of four terms,


corresponding to the multi-indices 1, 5 , 2, 3 , 2, 4 and 5, 3 , whereas consists of one term, corresponding to the multi-index 1, 6, 3, 2 .
Note, however, that could equally well be regarded as a 2-form on R 6 that
does not involve the variable x 6 . To avoid such ambiguities it is good practice to
state explicitly the domain of definition when writing a differential form.
Another reason for being precise about the domain of a form is that the coefficients f I may not be defined on all of Rn , but only on an open subset U of Rn . In
such a case we say is a k-form on U. Thus the expression ln x 2 y2 z dz is not
a 1-form on R3 , but on the open set U R3 x, y, z x2 y2 0 , i.e. the
complement of the z-axis.
17

18

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

You can think of dxi as an infinitesimal increment in the variable x i and of dx I


as the volume of an infinitesimal k-dimensional rectangular block with sides dx i1 ,
dxi2 , . . . , dx ik . (A precise definition will follow in Section 7.2.) By volume we here
mean oriented volume, which takes into account the order of the variables. Thus,
if we interchange two variables, the sign changes:
dxi1 dxi2 // dxiq / / dxi p / / dxik
dxi1 dxi2 // dxi p / / dxiq / / dxik ,

(2.1)

and so forth. This is called anticommutativity, graded commutativity, or the alternating property. In particular, this rule implies dx i dxi dxi dxi , so dxi dxi 0 for all
i.
Let us consider k-forms for some special values of k.
A 0-form on Rn is simply a smooth function (no dxs).
A general 1-form looks like
f 1 dx1

f 2 dx2

f n dxn .

//

A general 2-form has the shape

f i, j dxi dx j
i, j

f 1,1 dx1 dx1


f 2,1 dx2 dx1

f 1,2 dx1 dx2

f 1,n dx1 dxn

//

f 2,2 dx2 dx2 // f 2,n dx2 dxn //


f n,1 dxn dx1 f n,2 dxn dx2 // f n,n dxn dxn .

Because of the alternating property (2.1) the terms f i,i dxi dxi vanish, and a pair of
terms such as f 1,2 dx1 dx2 and f 2,1 dx2 dx1 can be grouped together: f 1,2 dx1 dx2
f 2,1 dx2 dx1 f 1,2 f 2,1 dx1 dx2 . So we can write any 2-form as

1 i j n

gi, j dxi dx j

g1,2 dx1 dx2

//

g1,n dx1 dxn

g2,3 dx2 dx3

/ / g2,n dx2 dxn


Written like this, a 2-form has at most
n n 1 n 2

// 2 1

// gn

1,n dx n 1 dx n .

1
n n 1
2

components.
Likewise, a general n 1-form can be written as a sum of n components,
f 1 dx2 dx3 // dxn

f 2 dx1 dx3 // dxn

//

f n dx1 dx2 // dxn

i 1

f i dx1 dx2 // dx
i // dxn ,

where dx
i means omit the factor dx i .
Every n-form on Rn can be written as f dx 1 dx2 // dxn . The special n-form
dx1 dx2 // dxn is also known as the volume form.
Forms of degree k n on Rn are always 0, because at least one variable has to
repeat in any expression dx i1 // dxik . By convention forms of negative degree are
0.
In general a form of degree k can be expressed as a sum

f I dx I ,
I

2.1. ELEMENTARY PROPERTIES

19

where the I are increasing multi-indices, 1 i 1 i2 Q// ik n. We shall


almost always represent forms in this manner. The maximum number of terms
occurring in is then the number of increasing multi-indices of degree k. An
increasing multi-index of degree k amounts to a choice of k numbers from among
the numbers 1, 2, . . . , n. The total number of increasing multi-indices of degree k
is therefore equal to the binomial coefficient n choose k,

n!
n
.
k k! n k !
(Compare this to the number of all multi-indices of degree k, which is n k .) Two
k-forms I f I dx I and I g I dx I (with I ranging over the increasing multi

indices of degree k) are considered equal if and only if f I


g I for all I. The

collection of all k-forms on an open set U is denoted by k U . Since k-forms can


be added together and multiplied by scalars, the collection k U constitutes a
vector space.
A form is constant if the coefficients f I are constant functions. The set of constant k-forms is a linear subspace of k U of dimension nk . A basis of this subspace is given by the forms dx I , where I ranges over all increasing multi-indices
of degree k. (The space k U itself is infinite-dimensional.)
The (exterior) product of a k-form
I f I dx I and an l-form
J g J dx J is

defined to be the k l-form

f I g J dx I dx J .

I,J

Usually many terms in a product cancel out or can be combined. For instance,

y2 x2 dx dy dz.

As an extreme example of such a cancellation, consider an arbitrary form of


degree k. Its p-th power p is of degree kp, which is greater than n if k 0 and
p n. Therefore
n 1 0

for any form on Rn of positive degree.


The alternating property combines with the multiplication rule to give the following result.
y dx x dy  x dx dz

y dy dz

y 2 dx dy dz x 2 dy dx dz

2.1. P ROPOSITION (graded commutativity).

for all k-forms and all l-forms .

1 kl

i1 , i2 , . . . , ik and J
P ROOF. Let I

the alternating property we get


dx I dx J

j1 , j2 , . . . , jl . Successively applying

dxi1 dxi2 // dxik dx j1 dx j2 dx j3 // dx jl

1 k dx j1 dxi1 dxi2 // dxik dx j2 dx j3 // dx jl

2k

dx j1 dx j2 dxi1 dxi2 // dxik dx j3 // dx jl

kl

dx J dx I .

..
.

20

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

For general forms

I f I dx I and

g J f I dx J dx I

J g J dx J we get from this


kl

I,J

f I g J dx I dx J
I,J

1 kl,

which establishes the result.

QED
2

A noteworthy special case is . Then we get 2


1 k 2 1 k 2 .
2
This equality is vacuous if k is even, but tells us that 0 if k is odd.
2.2. C OROLLARY. 2

0 if is a form of odd degree.


2.2. The exterior derivative

If f is a 0-form, that is a smooth function, we define d f to be the 1-form


n

xi dxi .

df

i 1

Then we have the product or Leibniz rule:


d f g $

f dg

g df .

If I f I dx I is a k-form, each of the coefficients f I is a smooth function and we


define d to be the k 1-form

d f I dx I .

The operation d is called exterior differentiation. An operator of this sort is called a


first-order partial differential operator, because it involves the first partial derivatives of the coefficients of a form.
f dx g dy is a 1-form on R2 , then

2.3. E XAMPLE . If
d

f y dy dx

g x dx dy g x

f y dx dy.

(Recall that f y is an alternative notation for f y.) More generally, for a 1-form
ni 1 f i dxi on Rn we have
n

fi
dx j dxi
x
j
i, j 1

d f i dxi

i 1

fi
dx j dxi
x
j
1 j i n
fj
f
q
xij dxi dx j xi dxi dx j
1 i j n
1 i j n
fj
f

xi xij dxi dx j ,
1 i j n

fi
dx j dxi
x
j
1 i j n

(2.2)

where in line (2.2) in the first sum we used the alternating property and in the
second sum we interchanged the roles of i and j.

2.2. THE EXTERIOR DERIVATIVE

f dx dy g dx dz h dy dz is a 2-form on R 3 , then

2.4. E XAMPLE . If
d

f z dz dx dy

g y dy dx dz h x dx dy dz f z g y h x dx dy dz.

For a general 2-form

1 i j n

1 i j
d f i, j dxi

n f i, j dx i dx j

on Rn we have

1 i j n k 1

f i, j
dxk dxi dx j
xk

f i, j
dxk dxi dx j
xk
1 k i j n

f i, j
dxk dxi dx j
xk
1 i k j n
f i, j

xk dxk dxi dx j
1 i j k n

f j,k
dxi dx j dxk
xi
1 i j k n

f i,k
dx j dxi dxk
x j
1 i j k n
f i, j

xk dxk dxi dx j
1 i j k n
f j,k
f i, j
f i,k

dxi dx j dxk .

xk
x j
xi
1 i j k n

21

(2.3)
(2.4)

Here in line (2.3) we rearranged the subscripts (for instance, in the first term we
relabelled k  i, i  j and j  k) and in line (2.4) we applied the alternating
property.
An obvious but quite useful remark is that if is an n-form on Rn , then d is
of degree n 1 and so d 0.
The operator d is linear and satisfies a generalized Leibniz rule.
2.5. P ROPOSITION .
(i) d a b a d b d for all k-forms and
and all scalars a and b.
(ii) d U d 1 k d for all k-forms and l-forms .
P ROOF. The linearity property (i) follows from the linearity of partial differentiation:
f
g
a f bg
a
b
xi
xi
xi
for al smooth functions f , g and constants a, b.
Now let I f I dx I and J g J dx J . The Leibniz rule for functions and
Proposition 2.1 give

I,J

I,J

f I g J dx I dx J

d f I dx I g J dx J

I,J

f I dg J g J d f I dx I dx J

1 k f I dx I dg J dx J 

1 k d,

which proves part (ii).


Here is one of the most curious properties of the exterior derivative.

QED

22

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

2.6. P ROPOSITION . d  d 

0 for any form . In short,

0.

d2
P ROOF. Let

I f I dx I . Then
n

fI
dx dx I
xi i
1

d

i 1

d  I dxi dx I .
xi 
I i
I i 1
Applying the formula of Example 2.3 (replacing f i with f I  xi ) we find
d  d 

fI
dxi
xi 

  

1 i j n

2 f I
xi x j

2 f I
dxi dx j
x j xi 

0,

because for any smooth (indeed, C 2 ) function f the mixed partials 2 f  xi x j and
2 f  x j xi are equal. Hence d  d  0.
QED
2.3. Closed and exact forms
A form is closed if d
one less).

0. It is exact if d for some form (of degree

2.7. P ROPOSITION . Every exact form is closed.


QED
d then d d  d 
0 by Proposition 2.6.
2.8. E XAMPLE . y dx  x dy is not closed and therefore cannot be exact. On
the other hand y dx  x dy is closed. It is also exact, because d  xy  y dx  x dy.
P ROOF. If

For a 0-form (function) f on Rn to be closed all its partial derivatives must vanish,
which means it is constant. A nonzero constant function is not exact, because
forms of degree 1 are 0.

Is every closed form of positive degree exact? This question has interesting
ramifications, which we shall explore in Chapters 4, 5 and 10. Amazingly, the
answer depends strongly on the topology, that is the qualitative shape, of the
domain of definition of the form.
Let us consider the simplest case of a 1-form ni 1 f i dxi . Determining
whether is exact means solving the equation dg for the function g. This
amounts to
g
g
g
f1 ,
f2 ,
fn,
...,
(2.5)
x1
x2
xn
a system of first-order partial differential equations. Finding a solution is sometimes
called integrating the system. By Proposition 2.7 this is not possible unless is
closed. By the formula in Example 2.3 is closed if and only if
fi
x j

fj
xi

for all 1  i  j  n. These identities must be satisfied for the system (2.5) to be
solvable and are therefore called the integrability conditions for the system.
2.9. E XAMPLE . Let
d

y dx  z cos yz  x  dy  y cos yz dz. Then

dy dx  z  y sin yz  cos yz  dz dy  dx dy



 y  z sin yz  cos yz  dy dz 0,


2.4. THE HODGE STAR OPERATOR

23

so is closed. Is exact? Let us solve the equations


g
x

y,

g
y

g
z

z cos yz  x,

y cos yz

by successive integration. The first equation gives g


yx  c y, z ! , where c is

a function of y and z only. Substituting into the second equation gives c " y

z cos yz, so c sin yz  k z ! . Substituting into the third equation gives k #
0, so


k is a constant. So g yx  sin yz is a solution and therefore is exact.

This method works always for a 1-form defined on all of Rn . (See Exercise 2.6.)
Hence every closed 1-form on Rn is exact.

$&% 0 ' defined by


$ y dx  x dy
y
x
$ 2
dx 
dy
.
2
2
2


x  y
x  y
x2  y2

2.10. E XAMPLE . The 1-form on R2

is called the angle form for reasons that will become clear in Section 4.3. From
x

x x2  y2

y2 $ x2
,
x2  y2 ! 2

y
y x2  y2

x2 $ y2
x2  y2 ! 2

it follows that the angle form is closed. This example is continued in Examples 4.1
and 4.6, where we shall see that this form is not exact.
For a 2-form
1 ( i ) j ( n f i, j dxi dx j and a 1-form

tion d amounts to the system

g j
xi

$ gi

x j

f i, j .

$ f i,k  f j,k
x j

xi

1 gi

dxi the equa(2.6)

By the formula in Example 2.4 the integrability condition d


f i, j
xk

ni*

0 comes down to

for all 1 + i , j , k + n. We shall learn how to solve the system (2.6), and its
higher-degree analogues, in Example 10.18.
2.4. The Hodge star operator
The binomial coefficient - nk . is the number of ways of selecting k (unordered)
objects from a collection of n objects. Equivalently, - nk . is the number of ways of
partitioning a pile of n objects into a pile of k objects and a pile of n $ k objects.
/
/
Thus we see that
n
n
.
k0 
n $ k0
This means that in a certain sense there are as many k-forms as n $ k-forms. In
fact, there is a natural way to turn k-forms into n $ k-forms. This is the Hodge star
operator. Hodge star of is denoted by 1 (or sometimes 2 ) and is defined as
follows. If I f I dx I , then

with

f I 31 dx I ! ,

I

1 dx I


I dx I c .

24

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

Here, for any increasing multi-index I, I c denotes the complementary increasing


multi-index, which consists of all numbers between 1 and n that do not occur in I.
The factor I is a sign,

1 if dx I dx I c 4 dx1 dx2 78787 dxn ,


465 9 1 if dx dx c 9 dx dx
I
I 4
1
2 78787 dx n .

In other words, : dx I is the product of all the dx j s that do not occur in dx I , times a
factor ; 1 which is chosen in such a way that dx I < : dx I = is the volume form:
dx I < : dx I =>4

dx1 dx2

78787 dxn .

2.11. E XAMPLE . Let n 4 6 and I 4?< 2, 6 = . Then I c


dx2 dx6 and dx I c 4 dx1 dx3 dx4 dx5 . Therefore

4@< 1, 3, 4, 5 = , so dx I 4

4 dx2 dx6 dx1 dx3 dx4 dx5


4 dx1 dx2 dx6 dx3 dx4 dx5 4 9 dx1 dx2 dx3 dx4 dx5 dx6 ,
which shows that I 4 9 1. Hence : < dx 2 dx6 =A4 9 dx1 dx3 dx4 dx5 .
2.12. E XAMPLE . On R2 we have : dx 4 dy and : dy 4 9 dx. On R3 we have
dx I dx I c

: dx 4 dy dz,
: dy 4 9 dx dz 4 dz dx,
: dz 4 dx dy,

: < dx dy =
4 dz,
: < dx dz =>4 9 dy,
: < dy dz =
4 dx.

(This is the reason that 2-forms on R3 are sometimes written as f dx dy B g dz dx B


h dy dz, in contravention of our usual rule to write the variables in increasing order.
In higher dimensions it is better to stick to the rule.) On R4 we have

: dx1 4 dx2 dx3 dx4 ,


: dx2 4 9 dx1 dx3 dx4 ,

: dx3 4 dx1 dx2 dx4 ,


: dx4 4 9 dx1 dx2 dx3 ,

and

: < dx1 dx2 =>4 dx3 dx4 ,


: < dx2 dx3 =A4 dx1 dx4 ,
9
: < dx1 dx3 =>4
: < dx2 dx4 =A4 9 dx1 dx3 ,
dx2 dx4 ,
: < dx1 dx4 =>4 dx2 dx3 ,
: < dx3 dx4 =A4 dx1 dx2 .
n
On R we have : 1 4 dx1 dx2 78787 dxn , : < dx1 dx2 78787 dxn =A4 1, and
for 1 G i G n,
: dxi 4C< 9 1 = i D 1dx1 dx2 78787Fdx
E i 78787 dxn
i
j
1
D
D
dx1 dx2 78787Hdx
for 1 G i J j G n.
: < dxi dx j =A4C< 9 1 =
E i 78787Idx
E j 78787 dxn
2.5. div, grad and curl
A vector field on an open subset U of Rn is a smooth map F : U
write F in components as
OQP
F1 < x = PP
F2 < x =
F < x =>4@LM
,
.
M .. R

MN

Fn < x =

Rn . We can

2.5. DIV, GRAD AND CURL

25

or alternatively as F S niT 1 Fi ei , where e1 , e2 , . . . , en are the standard basis vectors


of Rn . Vector fields in the plane can be plotted by placing the vector F U x V with
its tail at the point x. The diagrams below represent the vector fields W ye1 X xe2
and UW x X xy V e1 X U y W xy V e2 (which you may recognize from Exercise 1.10). The
arrows have been shortened so as not to clutter the pictures. The black dots are
the zeroes of the vector fields (i.e. points x where F U x V>S 0).
y

We can turn F into a 1-form by using the Fi as coefficients: S niT 1 Fi dxi .


For instance, the 1-form SYW y dx X x dy corresponds to the vector field F S
W ye1 X xe2 . Let us introduce the symbolic notation

]Q^^

dx

SYZ[[
[\

dx1
dx2
..
. _

^
,

dxn

which we will think of as a vector-valued 1-form. Then we can write S F ` dx.


Clearly, F is determined by and vice versa. Thus vector fields and 1-forms are
symbiotically associated to one another.
vector field F

acb

1-form :

S F ` dx.

Intuitively, the vector-valued 1-form dx represents an infinitesimal displacement.


If F represents a force field, such as gravity or an electric force acting on a particle,
then S F ` dx represents the work done by the force when the particle is displaced
by an amount dx. (If the particle travels along a path, the total work done by the
force is found by integrating along the path. We shall see how to do this in Section
4.1.)
The correspondence between vector fields and 1-forms behaves in an interesting way with respect to exterior differentiation and the Hodge star operator. For

26

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

d nie 1 f f g xi h dxi is associated to the vector field


mQn
f n
x 1 n
f nn
n
f
x 2
grad f d
e djik .
kk .. .
xi i
ie 1
kkl f o
x

each function f the 1-form d f

This vector field is called the gradient of f . (Equivalently, we can view grad f as
the transpose of the Jacobi matrix of f .)
grad f

pcq

df :

df

d grad f r dx.

Starting with a vector field F and letting

d F r dx, we find

s d n F s dx d n F 1 i v 1 dx dx r8r8rxdx
w i r8r8r dxn ,
1
2
i f i h i uf t h
ie 1
ie 1
Using the vector-valued n

t 1-form
mQnn
s dx mQnn
dx2 dx3 r8r8r dxn
1 n
n
s dx
dx
dx
dx
8
r
8
r
r
n
1
3
2
t
s dx d
..
kik ... o d ikk
o
.
v
n
1
lk s dxn
lk
dx1 dx2 r8r8r dxn y 1
fut 1 h
we can also write s d F r s dx. Intuitively, the vector-valued n t 1-form s dx
represents an infinitesimal n t 1-dimensional hypersurface perpendicular to dx.

(This point of view will be justified in Section 8.3, after the proof of Theorem 8.14.)
In fluid mechanics, the flow of a fluid or gas in Rn is represented by a vector field F.
The n t 1-form s then represents the flux, that is the amount of material passing
through the hypersurface s dx per unit time. (The total amount of fluid passing
through a hypersurface S is found by integrating over S. We shall see how to do
this in Section 5.1.) We have
ds

n
F
d d f F r s dx h d i ft 1 h i v 1 dxi dx1 dx2 r8r8rxdx
w i r8r8r dxn
xi
ie 1
n

d i dx1 dx2 r8r8r dxi r8r8r dxn d{z i dx1 dx2 r8r8r dxn .
xi
xi |
ie 1
ie 1
The function div F d nie 1 Fi g xi is the divergence of F. Thus if d F r dx, then
d s d d f F r s dx h d div F dx 1 dx2 r8r8r dxn .
An alternative way of writing this identity is obtained by applying s to both sides,

which gives

div F

d s d s .

A very different identity is found by first applying d and then


d

i, j

Fi
dx j dxi
x
j
1

} ~ } 

1 i j n

Fj
xi

s to :

i
t x j dxi dx j ,

EXERCISES

and hence

1 i j 1

1 i j nu

Fj
xi

27

Fi
dx1 dx2
x j

88Hdx
i 88dx
j 88 dxn .

In three dimensions d is a 1-form and so is associated to a vector field, namely

F3

curl F

x2
the curl of F. Thus, for n

F2
F3
F1
e1
e2
x3
x
x
1
3

3, if F dx, then

curl F dx

F2
F1
e3 ,
x
1 x 2

d .

You need not memorize every detail of this discussion. The point is rather to
remember that exterior differentiation in combination with the Hodge star unifies
and extends to arbitrary dimensions the classical differential operators of vector
calculus.
Exercises
2.1. Compute the exterior derivative of the following forms. Recall that a hat indicates
that a term has to be omitted.
(i) e xyz dx.
(ii) ni 1 x2i dx1 uudx
i uQ dxn .
n
p
i
(iii) x i 1 1 1 xi dx1 uu dx
i uQ dxn , where p is a real constant. For what values of p is this form closed?

2.2. Consider the forms


Calculate
(i) , ;
(ii) d , d, d .

x dx

y dy,

z dx dy

x dy dz and

z dy on R 3 .

2.3. Write the coordinates on R2n as x1 , y1 , x2 , y2 , . . . , xn , yn . Let

Compute d

dx1 dy1

dz x1 dy1

d d

u

dx2 dy2

uu

dxn dyn

x2 dy2

u

xn dyn

dxi dyi .
i 1
Compute n uu (n-fold product). First work out the cases n 1, 2, 3.
2.4. Write the coordinates on R2n 1 as x1 , y1 , x2 , y2 , . . . , xn , yn , z . Let

ye xy
3 4

2.6. Let
g x 

x1
0

z sin xz 3 dx
2xy z dx 3x2 y2 z4
ni

x3
0

3
R

xi dyi .

i 1

1, 2, 3.

is closed and find a function

xy
xey

z2 dy x sin xz F 2yz 3z 2 dz.


ze sin ze y 3 dy 4x2 y3 z3 e y sin ze y H e z dz.

f i dxi be a closed C

f 1 t, x2 , x3 , . . . , xn dt

dz

d . First work out the cases n

2.5. Check that each of the following forms


g such that dg .
(i)
(ii)

x2
0

1-form on Rn . Define a function g by

f 2 0, t, x3 , x4 , . . . , xn dt

f 3 0, 0, t, x4 , x5 , . . . , xn dt

Q

xn
0

f n 0, 0, . . . , 0, t dt.

Show that dg . (Apply the fundamental theorem of calculus, formula (B.3), differentiate
under the integral sign and dont forget to use d 0.)

28

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

2.7. Let ni 1 f i dxi be a closed 1-form whose coefficients f i are smooth functions
defined on Rn 0 that are all homogeneous of the same degree p 1. Let
1

g x
Show that dg

. (Use d

p 1

xi f i x .

i 1

0 and apply the identity proved in Exercise B.5 to each f i .)

2.8. Let and be closed forms. Prove that is also closed.


2.9. Let be closed and exact. Prove that is exact.
2.10. Calculate , , ,

, where , and are as in Exercise 2.2.


2.11. Consider the form x 22 dx1 x21 dx2 on R2 .
(i) Find and d d .
(ii) Repeat the calculation, regarding as a form on R 3 .
(iii) Again repeat the calculation, now regarding as a form on R 4 .

2.12. Prove that

1 kn

for every k-form on Rn .

2.13. Let I a I dx I and I b I dx I be constant k-forms, i.e. with constant coefficients a I and b I . (We also assume, as usual, that the multi-indices I are increasing.) The
inner product of and is the number defined by

, a I b I .
I

Prove the following assertions.


(i) The dx I form an orthonormal basis of the space of constant k-forms.
(ii) , 0 for all and , 0 if and only if 0.
(iii) , dx 1 dx2 uQ dxn .
(iv) .
(v) The Hodge star operator is orthogonal, i.e. , , .
2.14. The Laplacian of a smooth function on an open subset of R n is defined by

2 f
x21

2 f
x22

Q

2 f
.
x2n

Prove the following formulas.

(i) f d d f .

(ii) f g  f g f g 2 d f dg . (Use Exercise 2.13(iv).)


2.15.
(i)
(ii)
(iii)
2.16.

Let f : Rn R be a function and let f dx i .


Calculate d d .
Calculate d d .

Show that d d 1 n d d f dxi , where is the Laplacian defined in


Exercise 2.14.

(i) Let U be an open subset of Rn and let f : U R be a function satisfying grad f x 0 for all x in U. On U define a vector field n, an n 1-form
and a 1-form by
n x  grad f x Q

n dx,

grad f x Q

grad f x ,

df .

Prove that dx 1 dx2 uu dxn on U.


(ii) Let r : Rn R be the function r x x (distance to the origin). Deduce from
part (i) that dx 1 dx2 u dxn dr on Rn 0 , where  x 1 x dx.

EXERCISES

29

2.17. The Minkowski or relativistic inner product on R n

Rn

A vector x

x, y

xi yi

is spacelike if x, x

xn

i 1

is given by

1 yn 1 .

0, lightlike if x, x

0, and timelike if x, x

0.

(i) Give examples of (nonzero) vectors of each type.


(ii) Show that for every x 0 there is a y such that x, y 0.
A Hodge star operator corresponding to this inner product is defined as follows: if

I f I dx I , then

f I dx I ,

with

dx I

if I contains n 1,
if I does not contain n 1.

I dx I c
I dx I c

(Here I and I c are


Hodge star.)
as in the definition of the ordinary

(iii) Find 1, dx i for 1 i n 1, and dx 1 dx2 u dxn .


relativistic Laplacian (usually called the dAlembertian or wave
(iv) Compute the
f on Rn 1 .
operator) d d f for any smooth function

(v) For n 3 (ordinary space-time) find dx i dx j for 1 i j 4.

2.18. One of the greatest advances in theoretical physics of the nineteenth century was
Maxwells formulation of the equations of electromagnetism:
1 B
c t
4
1 D
J
c
c t
4

(Gau Law),

(no magnetic monopoles).

curl E

(Faradays Law),

curl H

(Ampres Law),

div D

div B

Here c is the speed of light, E is the electric field, H is the magnetic field, J is the density
of electric current, is the density of electric charge, B is the magnetic induction and D is
the dielectric displacement. E, H, J, B and D are vector fields and is a function on R 3 and
all depend on time t. The Maxwell equations look particularly simple in differential form

notation, as we shall now see. In space-time R4 with coordinates x 1 , x2 , x3 , x4 , where


x4 ct, introduce forms

E1 dx1

H1 dx1

E2 dx2

1
J dx dx
c 1 2 3

H2 dx2

E3 dx3 dx4

B1 dx2 dx3

H3 dx3 dx4

J2 dx3 dx1

B2 dx3 dx1

D1 dx2 dx3

J3 dx1 dx2 dx4

B3 dx1 dx2 ,

D2 dx3 dx1

D3 dx1 dx2 ,

dx1 dx2 dx3 .

(i) Show that Maxwells equations are equivalent to


d

d 4

0,
0.

(ii) Conclude that is closed and that div J t 0.


(iii) In vacuum one has E D and H B. Show that in vacuum
, the
relativistic Hodge star of defined in Exercise 2.17.

(iv) Free space is a vacuum without charges or currents.


Show that the Maxwell equations in free space are equivalent to d d 0.

30

2. DIFFERENTIAL FORMS ON EUCLIDEAN SPACE

(v) Let f , g : R

R be any smooth functions and define

E x f x1
g x1

x4
x4

B x 

0
g x1 x4
f x1 x4

Show that the corresponding 2-form satisfies the free Maxwell equations d
d 0. Such solutions are called electromagnetic waves. Explain why. In what
direction do these waves travel?

CHAPTER 3

Pulling back forms


3.1. Determinants
The determinant of a square matrix is the oriented volume of the block (parallelepiped) spanned by its column vectors. It is therefore not surprising that differential forms are closely related to determinants. This section is a review of some
fundamental facts concerning determinants. Let
A

...

a1,1
..
.

...

an,1

a1,n
..
.

an,n

be an n n-matrix with column vectors a1 , a2 , . . . , an . Its determinant is variously


denoted by
det A

det a1 , a2 , . . . , an det ai, j 1

i, j n

a1,1
6 ...

...

a1,n
.. .
.

an,1 . . . an,n

Expansion on the first column. You have probably seen the following definition of the determinant:
det A

 1 i 1 ai1 det Ai,1 .


i 1

Here Ai, j denotes the n 1 n 1 -matrix obtained from A by striking out the
i-th row and the j-th column. This is a recursive definition, which reduces the calculation of any determinant to that of determinants of smaller size. (The recursion
starts at n 1; the determinant of a 1 1-matrix a is simply defined to be the
number a.) It is a useful rule, but it has two serious flaws: first, it is extremely inefficient computationally (except for matrices containing lots of zeroes), and second,
it obscures the relationship with volumes of parallelepipeds.
Axioms. A far better definition is available. The determinant can be completely characterized by three simple laws, which make good sense in view of its
geometrical significance and which comprise an efficient algorithm for calculating
any determinant.
3.1. D EFINITION . A determinant is a function det which assigns to every ordered n-tuple of vectors a1 , a2 , . . . , an a number det a1 , a2 , . . . , an subject to the
following axioms:
31

32

3. PULLING BACK FORMS

(i) det is multilinear (i.e. linear in each column):


det a1 , a2 , . . . , cai

c ai , . . . , an
c det a1 , a2 , . . . , ai , . . . , an c det a1 , a2 , . . . , a , . . . , an
i
for all scalars c, c and all vectors a1 , a2 , . . . , ai , ai , . . . , an ;

(ii) det is alternating or antisymmetric:

det a1 , . . . , a j , . . . , ai , . . . , an

det a1 , . . . , ai , . . . , a j , . . . , an

for any i j;
(iii) normalization: det e1 , e2 , . . . , en
dard basis vectors of Rn .

1, where e1 , e2 , . . . , en are the stan-

We also write det A instead of det a1 , a2 , . . . , an , where A is the matrix whose


columns are a1 , a2 , . . . , an . Axiom (iii) lays down the value of det I. Axioms (i)
and (ii) govern the behaviour of oriented volumes under the elementary column
operations on matrices. Recall that these operations come in three types: adding a
multiple of any column of A to any other column (type I); multiplying a column
by a nonzero constant (type II); and interchanging any two columns (type III).
Type I does not affect the determinant, type II multiplies it by the corresponding
constant, and type III causes a sign change. This can be restated as follows.
3.2. L EMMA . If E is an elementary column operation, then det E A u
where
1 if E is of type I,

c if E is of type II (multiplication of a column by c),


k
1 if E is of type III.

k det A,

3.3. E XAMPLE . Identify the column operations applied at each step in the following calculation.
1
4
1

1
10
5

1
9
4

1
4
1

0
6
4

0
5
3

1
4
1

0
1
1

0
5
3

1
3
0

0
1
1

0
2
0

2 3 1 1 2 0 0 1 C 2 0 1 0 2.

As this example suggests, the axioms (i)(iii) suffice to calculate any n ndeterminant. In other words, there is at most one function det which obeys these
axioms. More precisely, we have the following result.
3.4. T HEOREM (uniqueness of determinants). Let det and det be two functions
satisfying Axioms (i)(iii). Then det A det A for all n n-matrices A.

P ROOF. Let a1 , a2 , . . . , an be the column vectors of A. Suppose first that A


is not invertible. Then the columns of A are linearly dependent. For simplicity
let us assume that the first column is a linear combination of the others: a1
c 2 a2
cn an . Applying axioms (i) and (ii) we get



det A

ci det ai , a2 , . . . , ai , . . . , an 0,

i 2

3.1. DETERMINANTS

 

33

and for the same reason det A


0, so det A
det A. Now assume that A
is invertible. Then A is column equivalent to the identity matrix, i.e. it can be
transformed to I by successive elementary column operations. Let E 1 , E2 , . . . , Em
E2 E1 A
I. According to
be these elementary operations, so that E m Em 1
Lemma 3.2, each operation E i has the effect of multiplying the determinant by a
certain factor k i , so axiom (iii) yields

  

   E E
A  k k    k k det A.
Applying the

same reasoning  to det A we get 1  k k    k k det A. Hence


det A  1  k k   k  det A.
QED

3.5. R
(change of normalization). Suppose that det is a function that
satisfies the multilinearity
 axiom (i) and the antisymmetry axiom (ii) but is normal
ized differently: det I  c. Then the proof of Theorem 3.4 shows that det A 
c det A for all n  n-matrices A.
1

det I

det Em Em

m m 1

2 1

2 1

m m 1

2 1

1 2

EMARK

This result leaves an open question. We can calculate the determinant of any
matrix by column reducing it to the identity matrix, but there are many different
ways of performing this reduction. Do different column reductions lead to the
same answer for the determinant? In other words, are the axioms (i)(iii) consistent? We will answer this question by displaying an explicit formula for the determinant of any n n-matrix that does not involve any column reductions. Unlike
Definition 3.1, this formula is not very practical for the purpose of calculating large
determinants, but it has other uses, notably in the theory of differential forms.

3.6. T HEOREM (existence of determinants). Every n


defined determinant. It is given by the formula
det A

Sn

n-matrix A has a well-

  a     a   .

sign a1,

2, 2

n, n

This requires a little explanation. S n stands for the collection of all permutations
of the set 1, 2, . . . , n . A permutation is a way of ordering the numbers 1, 2, . . . , n.
Permutations are usually written as row vectors containing each of these numbers
exactly once. Thus for n
2 there are only two permutations: 1, 2 and 2, 1 .
For n 3 all possible permutations are

1, 2, 3 ,

1, 3, 2 ,

2, 1, 3 ,

2, 3, 1 ,

3, 1, 2 ,
3, 2, 1 .

 1
n  2   3 2 1  n!
permutations. An alternative way of thinking of a permutation is as a bijective
(i.e. one-to-one and onto) map from the set  1, 2, . . . , n  to itself. For example, for
n  5 a possible permutation is

5, 3, 1, 2, 4 ,

and
as
a shorthand
notation for the map given
by 1  5,

2 we 3,think

3 of this

1, 4  2 and 5  4. The permutation 1, 2, 3, . . . , n  1, n


then corresponds to the identity map on the set  1, 2,
. . . , n  .

If is the identity permutation, then clearly i  j whenever i  j.


However, if is not the identity permutation, it cannot preserve the order in this
way. An inversion in is any pair of numbers i and j such that 1  i  j  n and
For general n there are

n n

34

3. PULLING BACK FORMS

   

 

i
j . The length of , denoted by l , is the number of inversions in .
A permutation is called even or odd according to whether its length is even, resp.
odd. For instance, the permutation 5, 3, 1, 2, 4 has length 6 and so is even. The
sign of is
1 if is even,
1 l
sign
1 if is odd.

$
"
#
    !   !
Thus sign  5, 3, 1, 2, 4 % 1. The permutations of & 1, 2 ' are  1, 2  , which has sign
1, and  2, 1  , which has sign ! 1, while for n  3 we have the table below.

l   sign  
 1, 2, 3  0
1
 1, 3, 2  1
!1
 2, 1, 3  1
!1
1
 2, 3, 1  2
1
 3, 1, 2  2
 3, 2, 1  3
!1
Thinking of permutations in S as bijective maps from & 1, 2, . . . , n ' to itself,
we can form the composition ( of any two permutations and in S . For
permutations we usually write instead of ( and call it the product of and
. This is the permutation produced by first performing and then ! For instance,
if   5, 3, 1, 2, 4  and   5, 4, 3, 2, 1  , then
  1, 3, 5, 4, 2  ,
) 4, 2, 1, 3, 5  .
n

A basic fact concerning signs, which we shall not prove here, is

 *

sign

 

 

sign sign .

(3.1)

In particular, the product of two even permutations is even and the product of an
even and an odd permutation is odd.
The determinant formula in Theorem 3.6 contains n! terms, one for each permutation . Each term is a product which contains exactly one entry from each
row and each column of A. For instance, for n
5 the permutation 5, 3, 1, 2, 4
contributes the term a 1,5 a2,3 a3,1 a4,2 a5,4 . For 2 2- and 3 3-determinants Theorem 3.6 gives the well-known formul

,, a

,, a
,, a
,, a

1,1
2,1
3,1

a1,2
a2,2
a3,2

a1,3
a2,3
a3,3

,,

,, a

,,

,, 

1,1
2,1

a1,2
a2,2

a1,1 a2,2 a3,3

,,

,, 

a1,1 a2,2

a1,1 a2,3 a3,2

a1,2 a2,1 ,

a1,2 a2,1 a3,3

a1,2 a2,3 a3,1

a1,3 a2,1 a3,2

a1,3 a2,2 a3,1 .

P ROOF OF T HEOREM 3.6. We need to check that the right-hand side of the
determinant formula in Theorem 3.6 obeys axioms (i)(iii) of Definition 3.1. Let
us for the moment denote the right-hand side by f A .
Axiom (i) is checked as follows: for every permutation the product
a1,

 
" # a " #/... a " #
1

2, 2

n, n

3.1. DETERMINANTS

35

contains exactly one entry from each row and each column in A. So if we multiply
the i-th row of A by c, each term in f A is multiplied by c. Therefore

0 1
f 0 a , a , . . . , ca , . . . , a 1*2
1

c f a1 , a2 , . . . , a i , . . . , a n .

Similarly,

3 a4 , . . . , a 1*2 f 0 a , a , . . . , a , . . . , a 1 3 f 0 a , a , . . . , a4 , . . . , a 1 .
Axiom (ii) holds because if we interchange two columns in A, each term in f 0 A 1
f a1 , a 2 , . . . , a i

changes sign. To see this, let be the permutation in S n that interchanges the two
numbers i and j and leaves all others fixed. Then

f a1 , . . . , a j , . . . , a i , . . . , a n

0 1

6 7 a 6 7/888 a 6 7
2 sign 0 1 a 6 7 a 6 7 888 a 6 7
substitute 2
5
2 sign 0 1 sign 0 1 a 6 7 a 6 7/888 a 6 7 by formula (3.1)
5
2:9 sign 0 1 a 6 7 a 6 7/888 a 6 7
by Exercise 3.4
5
2:9 f 0 a , . . . , a , . . . , a , . . . , a 1 .
Finally, rule (iii) is correct because if A 2 I,
1 if 2 identity,
a 6 7 a 6 7 888 a 6 7 2<;
0 otherwise,
and therefore f 0 I 12 1. So f satisfies all three axioms for determinants.
QED

Sn

sign a1,

2, 2

1, 1

Sn

n, n

2, 2

n, n

1, 1

Sn

1, 1

Sn
1

1, 1

2, 2

2, 2

2, 2

n, n

n, n

n, n

Here are some further rules followed by determinants. Each can be deduced
from Definition 3.1 or from Theorem 3.6. (Recall that the transpose of an n nai, j is the matrix A T whose i, j-th entry is a j,i.)
matrix A

2)0 1

3.7. T HEOREM . Let A and B be n

0 >1 2
2

n-matrices.

(i) det AB
det A det B.
(ii) det A T det A.
(iii) (Expansion on the j-th column) det A ni 1 1 i j ai, j det Ai, j for all j
1, 2, . . . , n. Here A i, j denotes the n 1
n 1 -matrix obtained from A
by striking out the i-th row and the j-th column.
(iv) det A a1,1 a2,2
an,n if A is upper triangular (i.e. a i, j 0 for i
j).

2 ? 0@9 1 A
0 9 1*=0 9 1

888

Volume change. We conclude this discussion with a slightly different geometric view of determinants. A square matrix A can be regarded as a linear map
A : Rn
Rn . The unit cube in Rn ,

D 0, 1E 2GF x H
n

I J

Rn 0

xi

for i

LD E M

1, 2, . . . , n ,

has n-dimensional volume 1. (For n


1 it is usually called the unit interval
and for n
2 the unit square.) Its image A 0, 1 n under the map A is a parallelepiped with edges Ae1 , Ae2 , . . . , Aen , the columns of A. Hence A 0, 1 n

LD E M

36

3. PULLING BACK FORMS

N O P SQ RUT TVR<T
vol A W X X RYT det A T vol X.

T O P

has n-dimensional volume vol A 0, 1 n


det A
det A vol 0, 1 n . This rule
n
generalizes as follows: if X is a measurable subset of R , then

e2

Ae2

AX

e1

Ae1

So det A can be interpreted as a volume change factor. (A set is measurable if it has


a well-defined, finite or infinite, n-dimensional volume. Explaining exactly what
this means is rather hard, but it suffices for our purposes to know that all open
and all closed subsets of Rn are measurable.)
3.2. Pulling back forms
By substituting new variables into a differential form we obtain a new form of
the same degree but possibly in a different number of variables.
3.8. E XAMPLE . In Example 2.10 we defined the angle form on R 2

Z\[ 0 ]

to be

Z y dx ^ x dy .
R
x ^ y
R
R
By substituting x cos t and y sin t into the angle form we obtain the following
1-form on R:
Z sin t d cos t ^ cos t d sin t R N W Z sin t X_W Z sin t X`^ cos tQ dt R dt.
cos t ^ sin t
2

We can take any k-form and substitute any number of variables into it to obtain
a new k-form. This works as follows. Suppose is a k-form defined on an open
subset V of Rm . Let us denote the coordinates on Rm by y1 , y2 , . . . , ym and let us
write, as usual,
f I dy I ,

where the functions f I are defined on V. Suppose we want to substitute new


variables x1 , x2 , . . . , xn and that the old variables are given in terms of the new by
functions

2 x1 , . . . , xn ,

y2

ym

1 x1 , . . . , xn ,

y1

..
.

m x1 , . . . , xn .

3.2. PULLING BACK FORMS

37

a b xc , where

As usual we write y

b c jh iii
b c .
b x caedf
ffg
k
b xc
1 x
2 x
..
.
m

We assume that the functions i are smooth and defined on a common domain U,
which is an open subset of Rn . We regard as a map from U to V. (In Example 3.8
we have U R, V R2
0 and t
cos t, sin t .) The pullback of along
is then the k-form on U obtained by substituting y i
i x1 , . . . , xn for all i
in the formula for . That is to say, is defined by

lnm o

p a

b c*a)b

b p

f I dy I .

p a

f I ,

Here f I is defined by

fI

c_b p c
q

p b ca b b cjc
a b c
a b
c
rrr dy c*a d d rrr d .

f I x ; in other words, f I
the composition of and f I . This means f I x
is the function resulting from f I by substituting y x . The pullback dy I is
defined by replacing each y i with i . That is to say, if I
i 1 , i2 , . . . , ik we put

dy I

a p b dy

i1

dyi2

i1

ik

i2

ik

The picture below is a schematic representation of the substitution process. The


form
I f I dy I is a k-form in y1 , y2 , . . . , ym ; its pullback
J g J dx J is a
k-form in x1 , x2 , . . . , xn . In Theorem 3.12 below we will give an explicit formula
for the coefficients g J in terms of f I and .

p a

t u xv
Rm

Rn
3.9. E XAMPLE . The formula

w xx x ayw ln b xx xz x c x
1
2

3
1 2
1
2

38

3. PULLING BACK FORMS


|

|Y} ~

defines a map : U
R2 , where U
x R2 x 1 x 2
0 . The components
3
of are given by 1 x1 , x2
x1 x2 and 2 x1 , x2
ln x1 x2 . Accordingly,

dy1

dy2

dy1 dy2

d1

d2
d1

| d x x | 3x x dx x dx ,
| d ln x x | x x dx dx ,
d | 3x x dx x dx _ x x dx
| 3xx x x x dx dx .
3
1 2

2
1 2

2
1 2

3
1

3
1

2
2 2

3
1

dx2

Observe that the pullback operation turns k-forms on the target space V into
k-forms on the source space U. Thus, while : U
V is a map from U to V, is
a map

V { U ,
k

the opposite way from what you might naively expect. (Recall that k U stands
for the collection of all k-forms on U.) The property that turns the arrow
around is called contravariance. Pulling back forms is nicely compatible with the
other operations that we learned about (except the Hodge star).

3.10. P ROPOSITION . Let : U


V be a smooth map, where U is open in R n and
m
V is open in R . The pullback operation is

|
|
| 

(i) linear: a b
a b ;
(ii) multiplicative:
;
(iii) natural:
, where : V
with W open in Rk and a form on W.

W is a second smooth map

The term natural in property (iii) is a mathematical catchword meaning that


a certain operation (in this case the pullback) is well-behaved with respect to composition of maps.

P ROOF. If
I f I dy I and I g I dy I are two forms of the same degree,
then a b I a f I bg I dy I , so

Now


a b |

 | a f

_ |

af

bg I dy I .


| a f x b g x ,
so a b | a f b g _ dy | a b . This proves part
(i). For the proof of part (ii) consider two forms | f dy and | g dy
(not necessarily of the same degree). Then | f g dy dy , so
| f g dy dy .
a fI

bg I x

bg I x

a fI x

bg I x

I,J I J

I J

I,J

Now

f g _ x |
I J

| f x g xj | f _ g _ x ,

fI gJ x

3.2. PULLING BACK FORMS

f g  f g . Furthermore,
dy dy * dy dy  dy dy  dy
d d  d d  d dy dy ,

so

I J

i1

i2

ik

i1

so

39

I,J

j1

jl

i2

ik

j1

jl


f dy @ g dy ,

f I g J dy I dy J
I

which establishes part (ii).


For the proof of property (iii) first consider a function f on W. Then

 f x  f x j f x
f x* f x ,
so f f . Next consider a 1-form dz on W, where z , z , . . . ,
dy , so
z are the variables on R . Then d


dy
d

y
y x dx y x dx .
By the chain rule, formula (B.6), the sum y x is equal to
x . Therefore

dx d  d  dz .

x
Because every form on W is a sum of products of forms of type f and dz , property
(iii) in general follows from the two special cases f and dz .
QED
f x

j 1

l 1

l 1

j 1

m
j 1

j 1

j 1

m i
j 1 y j

l 1

Another application of the chain rule yields the following important result.

3.11. T HEOREM . Let : U


V be a smooth map, where U is open in Rn and V is
m
k V . In short
open in R . Then d
d for


d d .

P ROOF. First let f be a function. Then

y f dy y f

df

i 1

i 1

di

y f x dx
y f
m

i 1

j 1

j 1i 1

i
dx j .
x j

40

3. PULLING BACK FORMS

By the chain rule, formula (B.6), the quantity m


i 1 f y i i x j is equal to
f x j . Hence

f
dx j
x j
1

df

d f ,

so the theorem is true for functions. Next let


so

I f I dy I . Then d

I d f I dy I ,

d f dy Y d f dy d f d d  d
because d f d f . On the other hand,
d d @ f dy @ d @ f d d  d
d f d d  d f d d d  d
d f d d  d .
d

i1

i2

ik ,

I
I
I

i1

i2

ik

i1

i2

ik

i1

i2

ik

i1

i2

ik

Here we have used the Leibniz rule for forms, Proposition 2.5(ii), plus the fact that
the form di1 di2
dik is always closed. (See Exercise 2.8.) Comparing the two
QED
equations above we see that d d .



We finish this section by giving an explicit formula for the pullback , which
establishes a connection between forms and determinants. Let us do this first in
degrees 1 and 2. The pullback of a 1-form m
i 1 f i dy i is


f dy f d .

Now di

i 1

nj

i
1 x j

i 1

dx j and so

f x dx g dx ,
n

i
dx j
x j
1

.
with g f
For a 2-form f dy dy we get

f dy dy f d d .


i 1

j 1i 1

j 1

i
i x j

m
i 1

1 i j m i, j

1 i j m

i, j

i, j

1 i j m

Observe that
di d j

k,l

i j
dxk dxl
xk xl
1

1 k l n

i j
xk xl

where

i j
xk xl

i j
xl xk

i
x k
j
x k

i
x l
j
x l

i j
dxk dxl ,
xl xk

3.2. PULLING BACK FORMS

41

is the determinant of the 2 2-submatrix obtained from the Jacobi matrix D by


extracting rows i and j and columns k and l. So we get

i
x k
f i, j
j
1 k l n x k

i
x k
j
x l

dx dx
>

f dx dx g dx dx

with

f
.
g

For an arbitrary k-form f dy we obtain

f dy dy  dy f d d  d .

To write the product d d  d in terms of the x-variables we use

dx
d
x
for l 1, 2, . . . , k. This gives


dx dx  dx
d d  d

x x x
x x  x dx ,
in which the summation is over all n multi-indices M
m , m , . . . , m . If a
multi-index M has repeating entries, then dx 0. If the entries of M are all

1 i j m

k,l

i
x k
j
x k

i, j

1 k l n1 i j m

i1

i1

ik

il

m 1 ,m 2 ,...,m k 1

k,l

i1

i2

ik

ml

ml

ml 1

ik

1 k l n

i
x l
j
x l

i2

i
x k
j
x k

ik

il

i1

i2

i2

i
x k
j
x l

i, j

1 i j m
I I

i1

i2

ik

m1

m2

mk

i1

i2

ik

m1

m2

mk

m1

m2

mk

distinct, we can rearrange them in increasing order by means of a permutation .


In other words, we have M
m1 , m2 , . . . , mk
j 1 , j 2 , . . . , j k , where
J
j1 , j2 , . . . , jk is an increasing multi-index and
S k is a permutation. Thus
we can rewrite the sum over all multi-indices M as a double sum over all increasing multi-indices J and all permutations :

di1 di1

 d

ik

i1 i2
x j 1 x j 2

J Sk

i1

dx
dx
dx

 x 
sign x x  x dx
(3.2)

ik

J Sk

i2

ik

det D I,J dx J .

(3.3)

In (3.2) used the result of Exercise 3.7 and in (3.3) we applied Theorem 3.6. The
notation D I,J stands for the I, J-submatrix of D, that is the k k-matrix obtained
from the Jacobi matrix by extracting rows i 1 , i2 , . . . , ik and columns j1 , j2 , . . . , jk .

42

3. PULLING BACK FORMS

To sum up, we find

f I det D I,J dx J
J

This proves the following result.

f I det D I,J

dx J .

3.12. T HEOREM . Let : U


V be a smooth map, where U is open in Rn and V is
m
open in R . Let
I f I dy I be a k-form on V. Then is the k-form on U given by
J g J dx J with

gJ

f I det D I,J .

This formula is seldom used to calculate pullbacks in practice and you dont
need to memorize the details of the proof. It is almost always easier to apply the
definition of pullback directly. However, the formula has some important theoretical uses, one of which we record here.
Assume that k m n, that is to say, the number of new variables is equal to
the number of old variables, and we are pulling back a form of top degree. Then

f det D dx dx  dx .
1 (constant function) then f 1, so we see that det D x can be in

  dy ,

f dy1 dy2

If f
terpreted as the ratio between the oriented volumes of two infinitesimal blocks
positioned at x: one with edges dx 1 , dx2 , . . . , dx n and another with edges d 1 ,
d2 , . . . , d n . Thus the Jacobi determinant is a measurement of how much the
map changes oriented volume from point to point.

3.13. T HEOREM . Let : U


V be a smooth map, where U and V are open in R n .
Then the pullback of the volume form on V is equal to the Jacobi determinant times the
volume form on U,

dy1 dy2

 dy det D dx dx  dx .
n

Exercises
3.1. Deduce Theorem 3.7(iv) from Theorem 3.6.

3.2. Calculate the following


Theorem 3.7(iv).
1
3
2
1
1
1
4
1

determinants using column and/or row operations and

1
5
2
3

1
2
,
3
7

2
3

1
1
1
1

2
1
1
2

4
3
.
0
5

3.3. Tabulate all permutations in S 4 with their lengths and signs.


3.4. Determine the length and the sign of the following permutations.
(i) A permutation of the form 1, 2, . . . , i 1, j, . . . , j 1, i, . . . , n where 1
i
j
n. (Such a permutation is called a transposition. It interchanges i and j and
leaves all other numbers fixed.)
(ii) n, n 1, n 2, . . . , 3, 2, 1 .

3.5. Find all permutations in S n of length 1.

EXERCISES

43

3.6. Calculate 1 , 1 , and , where


(i)
3, 6, 1, 2, 5, 4 and
5, 2, 4, 6, 3, 1 ;
(ii)
2, 1, 3, 4, 5, . . . , n 1, n and
n, 2, 3, . . . , n
positions interchanging 1 and 2, resp. 1 and n).

2, n

1, 1 (i.e. the trans-

3.7. Show that

dx sign dx dx dx
for any multi-index i , i , . . . , i and any permutation in S . (First show that the identity
is true if is a transposition. Then show it is true for an arbitrary permutation by writing
as a product j of transpositions and using formula (3.1) and Exercise 3.4(i).)
3.8. Show that for n 2 the permutation group S has n! 2 even permutations and
n! 2 odd permutations.
dxi 1 dxi 2
1

1 2

i1

i2

ik

3.9.

(i) Show that every permutation has the same length and sign as its inverse.
(ii) Deduce Theorem 3.7(ii) from Theorem 3.6.

1, 2, . . . , i 1, i 1, i, i
3.10. The i-th simple permutation is defined by i
So i interchanges i and i 1 and leaves all other numbers fixed. S n has n
permutations, namely 1 , 2 , . . . , n 1 . Prove the Coxeter relations

(i) i2 1 for 1 i n,
(ii) ii 1 3 1 for 1 i
(iii) i j 2 1 for 1 i, j

2, . . . , n .
1 simple

n 1,
n and i 1 j.
3.11. Let be a permutation of 1, 2, . . . , n . The permutation matrix corresponding to
is the n n-matrix A whose i-th column is the vector e . In other words, A e e .
(i) Write down the permutation matrices for all permutations in S .
(ii) Show that A A A .
(iii) Show that det A sign .
3.12.
(i) Suppose that A has the shape

a
a
... a

0
a
... a
A U .
,
..
.. ...
.
0
a
... a

1,1

1,2

1,n

2,2

2,n

n,1

n,n

i.e. all entries below a 11 are 0. Deduce from Theorem 3.6 that
det A

a1,1

...

a2,n
.. .
.
an,n

...

(ii) Deduce from this the expansion rule, Theorem 3.7(iii).

3.13. Show that

1
x1
x21
..
.

...
...
...

1
xn
x2n
..
.

x x

x
. . . x

from each row subtract x times the


for any numbers x , x , . . . , x . (Starting at the bottom,

row above it. This creates a new determinant whose first column is the standard basis vector
1

1
x2
x22
..
.

a2,2
..
.
an,2

n 1
1
n

n 1
2

n 1
n

i j

e1 . Expand on the first column and note that each column of the remaining determinant has
a common factor.)

44

3. PULLING BACK FORMS

x x . Find
x
x x
(i) dy , dy , dy ;
(ii) _ y y y , _ dy dy ;
(iii) dy dy dy .
x
x
x x

. Find
3.15. Let
x
xxx
(i) _ y
3y
3y
y ;
(ii) dy , dy , dy , dy ;
(iii) _ dy dy .
3.16. Compute _ x dy dz y dz dx

3.14. Let

x1

x1 x2

1 3

2 3

1 2 3
1

3
1
2
1 2
2
1 2 
3 
2

1 

2

3 

in Exercise B.7.

r cos sin

z dx dy , where is the map R2 




R3 defined

r cos cos

be spherical coordinates in R3 .

r sin
(i) Calculate P3 for the following forms :

3.17. Let P3

dx,

dy,

dz,

dx dy,

dx dz,

dy dz,

dx dy dz.

(ii) Find the inverse of the matrix DP3 .


as

3.18 (spherical coordinates in n dimensions). In this problem let us write a point in R n


r
1
..  .
. 




Let P1 be the function P1 r

Pn 

n 

r. For each n 

1 

1 define a map Pn 

.. 


n




n 

Rn 

1:

1 

Rn 

by

r
1
..   .
.  


cos P

n 

r sin n





(This is an example of a recursive definition. If you know P1 , you can compute P2 , and then
P3 , etc.)
(i) Show that P2 and P3 are the usual polar, resp. spherical coordinates on R 2 , resp.
R3 .
(ii) Give an explicit formula for P4 .
(iii) Let p be the first column vector of the Jacobi matrix of Pn . Show that Pn rp.
(iv) Show that the Jacobi matrix of Pn  1 is a n  1
n  1 -matrix of the form
DPn 

Av

u
,
w

where A is an n n-matrix, u is a column vector, v is a row vector and w is a


function given respectively by

sin P ,
w r cos .
v sin , 0, 0, . . . , 0 ,

cos n DPn ,
n

EXERCISES

45

(v) Show that det DPn 1  r cosn  1 n det DPn for n  1. (Expand det DPn 1 with
respect to the last row, using the formula in part (iv), and apply the result of part
(iii).)
(vi) Using the formula in part (v) calculate det DPn for n  1, 2, 3, 4.
(vii) Find an explicit formula for det DPn for general n.
(viii) Show that det DPn   0 for r   0.

CHAPTER 4

Integration of 1-forms
Like functions, forms can be integrated as well as differentiated. Differentiation and integration are related via a multivariable version of the fundamental
theorem of calculus, known as Stokes theorem. In this chapter we investigate the
case of 1-forms.
4.1. Definition and elementary properties of the integral
Let U be an open subset of Rn . A parametrized curve in U is a smooth mapping
c : I  U from an interval I into U. We want to integrate over I. To avoid problems
with improper integrals we assume I to be closed and bounded, I  a, b  . (Strictly
speaking we have not defined what we mean by a smooth map c :  a, b  U. The
easiest definition is that c should be the restriction of a smooth map c :  a  , b 
 U defined on a slightly larger open interval.) Let be a 1-form on U. The
pullback c  is a 1-form on  a, b  , and can therefore be written as c   g dt (where
t is the coordinate on R). The integral of over c is now defined by


 

a,b !

c


More explicitly, writing in components,


c

 c

so

i" 1


f i  dci

n 

i" 1 a

b
a

g  t  dt.

ni"


 c

i" 1

f i  c  t #

1 fi

fi 

dxi , we have

dci
dt,
dt

(4.1)

dci
 t  dt.
dt

4.1. E XAMPLE . Let U be the punctured plane R2  $ 0 % . Let c :  0, 2  U be


the usual parametrization of the circle, c  t &' cos t, sin t  , and let be the angle
form,
 y dx  x dy

.
x2  y2
Then c 


dt (see Example 3.8), so

2
( c )( 0

dt


2 .

A curve c :  a, b * U can be reparametrized by substituting a new variable,


t  p  s  , where s ranges over another interval  a , b  . We shall assume p to be a
Such
one-to-one mapping from  a , b  onto  a, b satisfying p +, s  - 0 for a . s . b.
a p is called a reparametrization. The parametrized curve
c / p :  a , b  U
has the same image as the original curve c, but it is traversed at a different rate.
Since p +  s 0 - 0 for all s 12 a , b  we have either p +  s &3 0 for all s (in which case p is
47

48

4. INTEGRATION OF 1-FORMS

increasing) or p 465 s 798 0 for all s (in which case p is decreasing). If p is increasing,
we say that it preserves the orientation of the curve (or that the curves c and c : p
have the same orientation); if p is decreasing, we say that it reverses the orientation
(or that c and c : p have opposite orientations). In the orientation-reversing case, c : p
traverses the curve in the opposite direction to c.
4.2. E XAMPLE . The curve c : ; 0, 2 <>= R2 defined by c 5 t 7@?A5 cos t, sin t 7 represents the unit circle in the plane, traversed at a constant rate (angular velocity)
of 1 radian per second. Let p 5 s 7B? 2s. Then p maps ; 0, < to ; 0, 2 < and c : p,
regarded as a map ; 0, <C= R2 , represents the same circle, but traversed at 2 radians per second. (It is important to restrict the domain of p to the interval ; 0, < .
If we allowed s to range over ; 0, 2 < , then 5 cos 2s, sin 2s 7 would traverse the circle
twice. This is not considered a reparametrization of the original curve c.) Now
let p 5 s 7D?FE s. Then c : p : ; 0, 2 <G= R2 traverses the unit circle in the clockwise
direction. This reparametrization reverses the orientation; the angular velocity is
now E 1 radian per second. Finally let p 5 s 7H? 2 s 2 . Then p maps ; 0, 1 < to ; 0, 2 <
and c : p : ; 0, 1 <I= R2 runs once counterclockwise through the unit circle, but at a
variable rate. What is the angular velocity as a function of s?
It turns out that the integral of a form along a curve is almost completely independent of the parametrization.
4.3. T HEOREM . Let be a 1-form on U and c : ; a, b <J=
p : ; a , b <=K; a, b < be a reparametrization. Then
L

cM p

?AN

if p preserves the orientation,


if p reverses the orientation.

c
O

c
O

U a curve in U. Let

P ROOF. It follows from the definition of the integral and from the naturality
of pullbacks (Proposition 3.10(iii)) that
L

LQP

cM p

Now let us write c S


5 p S g 7V5 dp W ds 7 ds, so

L P

a , b R

g dt and t
?

c : p 7#S

L P

cM p

On the other hand,


we have c M p ?YX
O

c ?
O
O

c ,

a , b R

pS g7

a , b R

p ST5 c S 7 .

p 5 s 7 . Then p S 5 c S 7&?
dp
ds
ds

L
?

b
a

p S 5 g dt 7U?A5 p S g 7 dp
?

g 5 p 5 s 7#7 p 4 5 s 7 ds.

b
a

g 5 t 7 dt, so by the substitution formula, Theorem B.7,


where the Z occurs if p 4\[ 0 and the E if p 4\8 0.
QED
O

Interpretation of the integral. Integrals of 1-forms play an important role


in physics and engineering. A curve c : ; a, b <*= U models a particle travelling
through the region U. Recall from Section 2.5 that to a 1-form ? ni] 1 Fi dxi corresponds a vector field F ? ni] 1 Fi ei , which can be thought of as a force field acting
on the particle. Symbolically we write ? F ^ dx, where we think of dx as an infinitesimal vector tangent to the curve. Thus represents the work done by the force
field along an infinitesimal vector dx. From (4.1) we see that c S ? F 5 c 5 t 7#7I^ c 4 5 t 7 dt.
Accordingly, the total work done by the force F on the particle during its trip along
c is the integral
L

F ^ dx
?

F 5 c 5 t 7_7`^ c 4 5 t 7 dt.

4.2. INTEGRATION OF EXACT 1-FORMS

49

In particular, the work and the total work are nil if the force is perpendicular to
the path, as in the picture on the left. The work done by the force in the picture on
the right is negative.

Theorem 4.3 can be translated into this language as follows: the work done by the
force does not depend on the rate at which the particle speeds along its path, but
only on the path itself and on the direction of travel.
The field F is conservative if it can be written as the gradient of a function,
F a grad g. The function b g is called a potential for the field and is interpreted as
the potential energy of the particle. In terms of forms this means that a dg, i.e.
is exact.
4.2. Integration of exact 1-forms
Integrating an exact 1-form

dg is easy once the function g is known.


a

4.4. T HEOREM (fundamental theorem of calculus in R n ). Let a dg be an exact


1-form on an open subset U of Rn . Let c : c a, b de U be a parametrized curve. Then
f

g g c g b h_h`b g g c g a h#h .

P ROOF. By Theorem 3.11 we have c i


g g c g t h#h we have c i a dh, so
a

fQk

a,b l

ci
a

c i dg
dh
a

dc i g. Writing h g t hja

c i g g t hja

h g b hmb h g a h ,

where we used the (ordinary) fundamental theorem of calculus, formula (B.1).


Hence n c a g g c g b h#hb g g c g a h#h .
QED
The physical interpretation of this result is that when a particle moves in a
conservative force field, its potential energy decreases by the amount of work done
by the field. This clarifies what it means for a field to be conservative: it means
that the work done is entirely converted into mechanical energy and that none is
dissipated by friction into heat, radiation, etc. Thus the fundamental theorem of
calculus explains the law of conservation of energy.

50

4. INTEGRATION OF 1-FORMS

It also yields a necessary and sufficient criterion for a 1-form on U to be exact.


A curve c : o a, b pq U is called closed if c r a sGt c r b s .
4.5. T HEOREM . Let be a 1-form on an open subset U of R n . Then the following
statements are equivalent.
(i) is exact.
(ii) u c t 0 for all closed curves c.
(iii) u c depends only on the endpoints of c for every curve c in U.
P ROOF. (i) tmv
(ii): if t dg and c is closed, then u c t g r c r b s#sIw g r c r a s_sxt
0 by the fundamental theorem of calculus, Theorem 4.4.
(ii) tv
(iii): assume u c t 0 for all closed curves c. Let
c1 : o a1 , b1 pq

and

c2 : o a2 , b2 pq

be two curves with the same endpoints, i.e. c 1 r a1 sDt c2 r a2 s and c1 r b1 s0t c2 r b2 s .
We need to show that u c1 tyu c2 . After reparametrizing c 1 and c2 we may
assume that a1 t a2 t 0 and b1 t b2 t 1. Define a new curve c by
c r t szt|{

for 0
for 1

c1 r t s
c2 r 2 w t s

t
t
}
}

1,
2.
}
}

(First traverse c 1 , then traverse c 2 backwards.) Then c is closed, so u c t 0. But


Theorem 4.3 implies u c t)u c1 w~u c2 , so u c1 t)u c2 .
(iii) tmv
(i): assume that, for all c, u c depends only on the endpoints of c.
We must define a function g such that t dg. Fix a point x 0 in U. For each point
x in U choose a curve cx : o 0, 1p`q U which joins x0 to x. Define
g r x sCt)

cx

We assert that dg is well-defined and equal to . Write t ni 1 f i dxi . We must


show that g xi t f i . From the definition of partial differentiation,
g
r x sjt
xi

lim

g r x hei s>w g r x s
h

1
0h

lim

cx

w~

hei

cx

Now consider a curve c composed of two pieces: for 0 } t } 1 travel from x0


to x along the curve cx and then for 1 } t } 2 travel from x to x hei along the
straight line given by l r t st x )r t w 1 s hei . Then c has the same endpoints as
cx hei . Therefore u cx he t)u c , and hence
i

g
r x sjt
xi

lim

1
h

w

cx

lim

1
h

cx
t

w

lim

1
h


cx

lim

1
h

1,2

l . (4.2)

4.3. THE GLOBAL ANGLE FUNCTION AND THE WINDING NUMBER

Let i, j be the Kronecker delta, which is defined by i,i


we can write l j t x j i, j t 1 h, and hence l j t
l

j 1

f j x t 1 hei dl j
n

j 1

j 1

f j x t

1 and i, j 0 if i j. Then
i, j h. This shows that

f j x t 1 hei l j t dt

1 hei i, jh dt

51

h f i x t 1 hei dt. (4.3)

Taking equations (4.2) and (4.3) together we find


g
x
xi

1
0h

lim

2
1

h f i x t 1 hei dt
1

lim f i x shei ds

0 h

This proves that g is smooth and that dg

lim

0 0

f i x ds

f i x shei ds

fi x .

QED

This theorem, and its proof, can be used in many different ways. For example,
it tells us that once we know a 1-form to be exact we can find an antiderivative
g x by integrating along an arbitrary path running from a fixed point x0 to x.
(See Exercises 4.34.5 for an application.) On the other hand, the theorem also
enables us to detect closed 1-forms that are not exact.
4.6. E XAMPLE . The angle form 1 R2 0 of Example 2.10 is closed,
but not exact. Indeed, its integral around the circle is 2 0. Mark the contrast with closed 1-forms on Rn , which are always exact! (See Exercise 2.6.) This
phenomenon underlines the importance of being careful about the domain of definition of a form.
4.3. The global angle function and the winding number
In this section we will have a closer look at the angle form and see that it
carries interesting information of a topological nature. Throughout this section
U will be the punctured plane R2 0 , will denote the angle form,

y dx x dy
,
x2 y2

and and will denote the functions


x

,
x2 y2

y
x2

y2

Then is a closed 1-form and and are smooth functions on U. In fact, and
are just the components of x \ x , the unit vector pointing in the direction of x.
These functions satisfy
d d .
(4.4)
(You will be asked to check this formula in Exercise 4.6.) Now let : U 0, 2
be the angle between a point and the positive x-axis, chosen to lie in the interval
0, 2 . Then cos and sin , so by equation (4.4)

cos d sin

sin d cos

cos 2 d

sin 2 d

d .

This equation is not valid on all of U (it cannot be because we saw in Example 4.6
that is not exact), but only where is differentiable, i.e. on the complement of the

52

4. INTEGRATION OF 1-FORMS

positive x-axis. Hence the nonexactness of is closely related to the impossibility


of defining a global differentiable angle function on U. (The precise meaning of
this assertion will become clear in Exercise 4.6.)
However, along a curve c : a, b
U we can define a continuous angle function, and the fact that d almost everywhere suggests how: by integrating
along c! For simplicity assume that a 0 and b 1. Start by fixing any 0 such
that cos 0 c 0 # and sin 0 c 0 # and then define

t j

0,t

c .

The following result says that t measures the angle between c t and the positive x-axis (up to an integer multiple of 2 ) and that the function : 0, 1
R is
smooth. In this sense is a differentiable choice of angle along the curve c.
4.7. T HEOREM . The function is smooth and satisfies

0 j

0 ,

cos t j c t #

and

sin t z

c t # .

P ROOF. To see that 0 * 0 , plug t 0 into the definition of . To prove


the other assertions we rescale the curve c t to a new curve c t #\ c t moving
on the unit circle. Let f t and g t be the x- and y-components of this new curve.
Then f t j c t _ and g t z c t # and
2

f t

g t

for all t. In other words f c and g


c f dg g d f . Therefore

t j

c , so it follows from formula (4.4) that

f s g s > g s f 6 s ds.

By the fundamental theorem of calculus, formula (B.2), is differentiable and

f g

gf .

(4.5)

Since the right-hand side is smooth, is smooth as well. To prove that cos t @
f t and sin t j g t for all t it is enough to show that the difference vector

f t
g t

cos t
sin t _

has length 0. Its length is equal to

cos

g sin

f2

g2

2 f cos g sin

Hence we need to show that the function u


to 1. For t 0 we have
u 0 j

f 0 cos 0

f cos

g 0 sin 0

cos 2 sin 2

2 f cos

g sin .

g sin is a constant equal

cos 2 0 sin 2 0

1.

Furthermore, the derivative of u is


f cos

f2f

f f f

f sin

g f

f gg cos

f gg cos

g sin g cos

gg cos

g2 g

g gg

f 2 g

f g f sin

f g f sin
f f sin .

by formula (4.5)
since f 2

g2

EXERCISES

Now f 2 g2 1 implies f f gg
function, so u t 1 for all t.

53

0 for all t. Hence u is a constant


QED

c t #\ c t T as a dial that points

0, so u t

It is useful to think of the vector f t , g t #


in the same direction as the vector c t .

As t increases from 0 to 1, the dial starts at the angle 0 0 , it moves around


the meter, and ends up at the final angle 1 . The difference 1 I 0 measures
the total angle swept out by the dial.
4.8. C OROLLARY. If c : 0, 1
where k is an integer.
P ROOF. By Theorem 4.7, c 0

cos 0 , sin 0

U is a closed curve, then 1 z 0

2 k,

c 1 implies

c 0 # , c 0 ##

c 1 # , c 1 ##

In other words cos 0 cos 1 and sin 0


by an integer multiple of 2 .

cos 1 , sin 1 .

sin 1 , so 0 and 1 differ


QED

The integer k 2 _ 1 c is called the winding number of the closed curve c


about the origin. It measures how many times the curve loops around the origin.
winding number of a closed curve about origin

1
2

4.9. E XAMPLE . By Example 4.1, the winding number of the circle c t


(0 t 2 ) is equal to 1.

(4.6)

cos t, sin t

Exercises
4.1. Consider the curve c : 0, 2 \ R2 defined by c t G) a cos t, b sin t T , where a
and b are positive constants. Let xy dx x 2 y dy.
(i) Sketch the curve c for a 2 and b
(ii) Find c (for arbitrary a and b).

1.

4.2. Restate Theorem 4.5 in terms of force fields, potentials and energy. Explain why
the result is plausible on physical grounds.

54

4. INTEGRATION OF 1-FORMS

4.3. Consider the 1-form ) x a ni 1 xi dxi on Rn J 0 , where a is a real constant.


For every x 0 let cx be the line segment starting at the origin and ending at x.
(i) Show that is closed for any value of a.
(ii) Determine for which values of a the function g x 9 cx is well-defined and
compute it.
(iii) For the values of a you found in part (ii) check that dg .
4.4. Let 1 Rn 0 be the 1-form of Exercise 4.3. Now let c x be the halfline
pointing from x radially outward to infinity. Parametrize c x by travelling from infinity
inward to x. (You can do this by using an infinite time interval U , 0 in such a way that
cx 0  x.)
(i) Determine for which values of a the function g x 9 cx is well-defined and
compute it.
(ii) For the values of a you found in part (i) check that dg .
(iii) Show how to recover from this computation the potential energy for Newtons
gravitational force. (See Exercise B.4.)
4.5. Let 1 Rn 0 be as in Exercise 4.3. There is one value of a which is not
covered by Exercises 4.3 and 4.4. For this value of a find a smooth function g on R n J 0
such that dg .
4.6.
(i) Verify equation (4.4).
(ii) Let U R2 0 . Prove that there does not exist a smooth function : U
R satisfying cos x, y x x2 y2 and sin x, y y x2 y2 for all
x, y U. (Argue by contradiction, by letting y dx x dy x 2 y 2
and showing that d if was such a function.)
4.7. Calculate directly from the definition the winding number about the origin of the
curve c : 0, 2 R2 given by c t m cos kt, sin kt T .
4.8. Let x0 be a point in R2 and c a closed curve which does not pass through x 0 .
How would you define the winding number of c around x 0 ? Try to formulate two different
definitions: a geometric definition and a definition in terms of an integral over c of a
certain 1-form analogous to formula (4.6).
4.9. Let c : 0, 1 Q R2 B 0 be a closed curve with winding number k. Determine the
winding numbers of the following curves c : 0, 1 m R2 0 by using the formula, and
then explain the answer by appealing to geometric intuition.
t  c 1 t ;
t  t c t , where : 0, 1 Q 0, is a function satisfying 0 
t  c t _ 1 c t ;
t  c t , where x, y > y, x T ;
T
1
(v) c t  c t , where x, y >
x, y .
2
2

x
y

(i)
(ii)
(iii)
(iv)

c
c
c
c

1 ;

4.10. For each of the following closed curves c : 0, 2 Q R2 0 set up the integral
defining the winding number about the origin. Evaluate the integral if you can (but dont
give up too soon). If not, sketch the curve (the use of software is allowed) and obtain the
answer geometrically.
(i)
(ii)
(iii)
(iv)

c
c
c
c

t  a cos t, b sin t T , where a 0 and b 0;


t  cos t 2, sin t T ;
t m cos3 t, sin3 t T ;
t ) a cos t b cos t b a 2, a cos t b sin t

, where 0

 b  a.

EXERCISES

R2

4.11. Let b

0  by

c  t

55

 0 and a   0 be constants with  a   b. Define a planar curve c :  0, 2 



  a  b  cos t  a cos a  b t,  a  b  sin t  a sin a  b t 
a

(i) Sketch the curve c for a  b  3.


(ii) For what values of a and b is the curve closed?
(iii) Assume c is closed. Set up the integral defining the winding number of c around
the origin and evaluate it. If you get stuck, find the answer geometrically.

 F1 e1  F2 e2 : U

4.12. Let U be an open subset of R2 and let F


vector field. The differential form

R2 be a smooth


 F1 dF2 F2 dF1
F12  F22

is well-defined at all points x of U where F  x   0. Let c be a parametrized circle contained


in U, traversed once in the counterclockwise direction. Assume that F  x   0 for all x  c.
The index of F relative to c is
index  F, c 

1
2

Prove the following assertions.


(i)
(ii)
(iii)
(iv)

4.13.

 F  , where is the angle form  y dx  x dy   x2  y2 ! ;


is closed;
index  F, c  is the winding number of the curve F " c about the origin;
index  F, c  is an integer.

circles.

(i) Find the indices of the following vector fields around the indicated

56

4. INTEGRATION OF 1-FORMS

(ii) Draw diagrams of three vector fields in the plane with respective indices 0, 2
and 4 around suitable circles.

CHAPTER 5

Integration and Stokes theorem


5.1. Integration of forms over chains
In this chapter we generalize the theory of Chapter 4 to higher dimensions. In
the same way that 1-forms are integrated over parametrized curves, k-forms can
be integrated over k-dimensional parametrized regions. Let U be an open subset
of Rn and let be a k-form on U. The simplest k-dimensional analogue of an
interval is a rectangular block in Rk whose edges are parallel to the coordinate axes.
This is a set of the form
R #%$ a1 , b1 &(' $ a2 , b2 &('*)+)+)' $ ak , bk & #, t - Rk . ai /

ti /

bi

for 1 /

i/

k0 ,

where ai 1 bi . The k-dimensional analogue of a parametrized path is a smooth


map c : R 2 U. Although the image c 3 R 4 may look very different from the block
R, we think of the map c as a parametrization of the subset c 3 R 4 of U: each choice
of a point t in R gives rise to a point c 3 t 4 in c 3 R 4 . The pullback c 5 is a k-form
on R and therefore looks like g 3 t 4 dt 1 dt2 )+)+) dtk for some function g : R 2 R. The
integral of over c is defined as

c5 #

bk
ak

)+)+)

b2
a2

b1
a1

g 3 t 4 dt1 dt2 )+)+) dtk .

For k # 1 this reproduces the definition given in Chapter 4. (The definition makes
sense if we replace the rectangular block R by more general shapes in R k , such as
skew blocks, k-dimensional balls, cylinders, etc. In fact any compact subset of R k
will do.)
The case k # 0 is also worth examining. A zero-dimensional block R in
R0 #7, 0 0 is just the point 0. We can therefore think of a map c : R 2
U as a
collection , x 0 consisting of a single point x # c 3 0 4 in U. The integral of a 0-form
(function) f over c is by definition the value of f at x,

f #

f 3 x4 .

As in the one-dimensional case, integrals of k-forms are almost wholly unaffected by a change of variables. Let
R #8$ a 1 , b 1 & ' $ a 2 , b 2 & '*)+)+)9' $ a k , b k &
be a second rectangular block. A reparametrization is a map p : R 2 R satisfying
the following conditions: p is bijective (i.e. one-to-one and onto) and the k ' k Then det Dp 3 s 4;# : 0 for all s - R,
so either
matrix Dp 3 s4 is invertible for all s - R.
det Dp 3 s 4=< 0 for all s or det Dp 3 s 4 1 0 for all s. In these cases we say that the
reparametrizion preserves, respectively reverses the orientation of c.
57

58

5. INTEGRATION AND STOKES THEOREM

U a smooth map. Let p : R >

5.1. T HEOREM . Let be a k-form on U and c : R >


be a reparametrization. Then

ACB E D c
c@ p
c

if p preserves the orientation,


if p reverses the orientation.

P ROOF. Almost verbatim the same proof as for k A 1 (Theorem 4.3). It follows
from the definition of the integral and from the naturality of pullbacks, Proposition
3.10(iii), that
?
?
?
c@ p

Now let us write c I A


p I c I HLA

R F

c G p HJI A

g dt1 dt2 K+K+K dtk and t A


p I g dt1 dt2 K+K+K dtk HLA

by Theorem 3.13, so

pI cI H .

p s H . Then

p I g H det Dp ds1 ds2 K+K+K dsk

g p sHJH det Dp s H ds 1 ds2 K+K+K dsk .

F F
F
On the other hand, c A
g t H dt1 dt2 K+K+K dtk , so by the substitution formula,
D
DR
Theorem B.7, we have c @ p ANM F c , where the O occurs if det Dp P 0 and the
D
D
E if det Dp Q 0.
QED
c@ p

5.2. E XAMPLE . The unit interval is the interval R 0, 1 S in the real line. Any curve
c : R a, bST>
U can be reparametrized to a curve c G p : R 0, 1 SU>
U by means of
the reparametrization p s HVA
b E a H s O a. Similarly, the unit cube in R k is the
F
F
rectangular block

R 0, 1 S k AXW t Y Rk Z ti Y[R 0, 1 S
Let R be any other block, given by a i \
As O a, where
b1 E a1
0
A_
A ^`
..
`
.

`a

0
b2 E a2
..
.
0

...
...
..
.
...

ti \

for 1 \

i\

k] .

bi . Define p : R 0, 1 S

bdc
c

bk E ak

>

R by p s H;A

bdc

0
0
..
.

and

a1 cc
a2
a A7^` . .
` .. e

`a

ak

(Squeeze the unit cube until it has the same edgelengths as R and then move it to
the position of R.) Then p is one-to-one and onto and Dp sHfA A, so det Dp s HLA
F
F
det A A vol R P 0 for all s, so p is an orientation-preserving
reparametrization.
Hence c @ p A
c for any k-form on U.

5.3. R EMARK . A useful fact you learned in calculus is that one may interchange the order of integration in a multiple integral, as in the formula

b1
a1

b2
a2

f t1 , t2 H dt1 dt2 A

b2
a2

b1
a1

f t1 , t2 H dt2 dt1 .

(5.1)

(This follows for instance from the substitution formula, Theorem B.7.) On the
other hand, we have also learned that f t 1 , t2 H dt2 dt1 A E f t1 , t2 H dt1 dt2 . How
F follows. Let A
can this be squared with formula (5.1)? F The explanation is as
f t1 , t2 H dt1 dt2 . Then the left-hand side of formula (5.1) is the integral of over

5.2. THE BOUNDARY OF A CHAIN

59

c : g a1 , b1 h i g a2 , b2 hkj R2 , the parametrization of the rectangle given by c l t 1 , t2 mon


l t1 , t2 m . The right-hand side is the integral of p not over c, but over c q p, where

p : g a2 , b2 h(i g a1 , b1 h j g a1 , b1 h i g a2 , b2 h
is the reparametrization p l s 1 , s2 mLn l s2 , s1 m . Since p reverses the orientation, Theorem 5.1 says that r c s p n ptr c ; in other words r c n r c s p lup m , which is exactly
formula (5.1). Analogously we have

vxw

0,1 y k

f l t1 , t2 , . . . , tk m dt1 dt2 z+z+z dtk n

v
w

0,1 y k

f l t1 , t2 , . . . , tk m dti dt1 dt2 z+z+z|d{ ti z+z+z dtk

for any i.
We see from Example 5.2 that an integral over any rectangular block can be
written as an integral over the unit cube. For this reason, from now on we shall
usually take R to be the unit cube. A smooth map c : g 0, 1 h k j U is called a k-cube
in U (or sometimes a singular k-cube, the word singular meaning that the map c is
not assumed to be one-to-one, so that the image can have self-intersections.)
It is often necessary to integrate over regions that are made up of several
pieces. A k-chain in U is a formal linear combination of k-cubes,
c n a1 c1 } a2 c2 }~z+z+z|} a p c p ,
where a1 , a2 , . . . , a p are real coefficients and c 1 , c2 , . . . , c p are k-cubes. For any
k-form we then define

ai

i 1

ci

(In the language of linear algebra, the k-chains form an abstract vector space with
a basis consisting of the k-cubes. Integration, which is a priori only defined on
cubes, is extended to chains in such a way as to be linear.)
Recall that a 0-cube is nothing but a singleton x consisting of a single point
p
x in U. Thus a 0-chain is a formal linear combination of points, c n i  1 ai xi .
A good way to think of c is as a collection of p point charges, with an electric
charge ai placed at the point xi . (You must carefully distinguish between the formal
p
linear combination i  1 ai xi , which represents a distribution of point charges,
p
and the linear combination of vectors i  1 ai xi , which represents a vector in Rn .)
The integral of a function f over the 0-chain is by definition

c
p

f n

i 1

ai

xi

f n

a i f l xi m .

i 1

Likewise, a k-chain i  1 ai ci can be pictured as a charge distribution, with an electric charge ai spread along the k-dimensional patch c i .
5.2. The boundary of a chain
Consider a curve (1-cube) c : g 0, 1 hj
0-chain defined by c n c l 1 m p[ c l 0 m .

c 0 d

U. Its boundary is by definition the

c 1 d

60

5. INTEGRATION AND STOKES THEOREM

The boundary of a 2-cube c : 0, 1 2 U consists of four pieces corresponding


to the edges of the unit square: c 1 t c t, 0 , c2 t c 1, t , c3 t c t, 1 and
c4 t c 0, t . The picture below suggests that we should define c c 1 c2
c3 c4 .

(Alternatively we could define c c 1 c2 c3 c4 , with c3 t 


c4 t o c 0, 1 t , which corresponds to the following picture:

c 1 t, 1 and

This would work equally well, but is technically less convenient.)


A k-cube c : 0, 1 k U has 2k faces of dimension k 1, which are described
as follows. Let t t1 , t2 , . . . , tk 1 [ 0, 1 k 1 and for i 1, 2, . . . , k put
ci,0 t f

ci,1 t f

c t1 , t2 , . . . , ti
c t1 , t2 , . . . , ti

1 , 0, t i , . . . , t k 1
1 , 1, t i , . . . , t k 1

,
.

(Insert 0, resp. 1 in the i-th slot.) Now define


c

i 1

1 i ci,0 ci,1 f

i 1 0,1

i
ci, .
u 1

For an arbitrary k-chain c i ai ci we put c i ai ci . Then is a linear map


from k-chains to k 1-chains. You should check that for k 0 and k 1 this
definition is consistent with the one- and two-dimensional cases considered above.
There are a number of curious similarities between the boundary operator
and the exterior derivative d, the most important of which is the following. (There
are also many differences, such as the fact that d raises the degree of a form by 1,
whereas lowers the dimension of a chain by 1.)
5.4. P ROPOSITION . c L

0 for every k-chain c in U. In short,


2

0.

P ROOF. By linearity of it suffices to prove this for k-cubes c : 0, 1 k U. Let


t1 , t2 , . . . , tk 2  0, 1 k 2 and let and be 0 or 1. Then for 1 i j k 1

5.3. CYCLES AND BOUNDARIES

we have

ci,

j,

61

t1 , t2 , . . . , tk

ci, t1 , t2 , . . . , t j

2 f

1, , t j , . . . , tk 2

c t1 , t2 , . . . , ti 1 , , ti , . . . , t j 1 , , t j , . . . , tk 2 .
other hand,

On the
c j 1, i, t1 , t2 , . . . , tk 2 L c j 1, t1 , t2 , . . . , ti 1 , , ti , . . . , tk 2

c t1 , t2 , . . . , ti 1 , , ti , . . . , t j 1, , t j , . . . , tk 2 ,

because in the vector t 1 , t2 , . . . , ti 1 , , ti , . . . , tk 2 the entry t j occupies the j 1st

slot! We conclude that c i, j,


c j 1, i, for 1 i j k 1. Therefore the
k 2-chain c is given by

k
k
c o 1 i ci, 1 i ci,
i 1 0,1
i 1 0,1

k k 1
i j
ci, j, .
1
i 1 j 1 , 0,1
The double sum over i and j can be rearranged in a sum over i
i j to give

c L

1 i j k 1 , 0,1

1 i j

ci,

j,

1 j i k , 0,1

1 i j

In the first term on the right in (5.2) we substitute c i, j,


r j 1, s i, , to get

1 i j k 1 , 0,1

1 i j

ci,

j,

1 i j k 1 , 0,1

1 s r k , 0,1

j and a sum over

c j

j, .

ci,

1, i,

1 i j

(5.2)

and then

c j

1, i,

1 s r 1 cr,

s r
1
cr,

1 s r k , 0,1

Thus the two terms on the right in (5.2) cancel out.

s,
s, .

QED

5.3. Cycles and boundaries


A k-cube c is degenerate if c t 1 , . . . , tk is independent of ti for some i. A k-chain
c is degenerate if it is a linear combination of degenerate cubes. In particular, a degenerate 1-cube is a constant curve. The work done by a force field on a motionless
particle is 0. More generally we have the following.
5.5. L EMMA . Let be a k-form and c a degenerate k-chain. Then c

0.

P ROOF. By linearity we may assume that c is a degenerate cube. Suppose c is


constant as
a function of ti . Then

c t1 , . . . , ti , . . . , tk f

c t1 , . . . , 0, . . . , t k f

g f t1 , . . . , ti , . . . , tk J ,

62

5. INTEGRATION AND STOKES THEOREM

where f : 0, 1

0, 1 k

and g : 0, 1 k

U are given respectively by

t1 , . . . , ti , . . . , tk ,
g s1 , . . . , sk 1 f c s1 , . . . , si 1 , 0, si 1, . . . , tk 1 .
Now g is a k-form on 0, 1 k 1 and hence equal to 0, and so c f g  0.
We conclude c + 0,1 k c 0.
QED
f t1 , . . . , ti , . . . , tk f

So degenerate chains are irrelevant where integration is concerned. This motivates the following definition. A k-chain c is closed, or a cycle, if c is a degenerate
k 1-chain. A k-chain c is a boundary if c b c for some k 1-chain b and
some degenerate k-chain c .
5.6. E XAMPLE . If c 1 and c2 are curves arranged head to tail as in the picture
below, then c1 c2 is a 1-cycle. Likewise, the closed curve c is a 1-cycle.
c2
c
c1
5.7. L EMMA . The boundary of a degenerate k-chain is a degenerate k 1-chain.
P ROOF. By linearity it suffices to consider the case of a degenerate k-cube c.
Suppose c is constant as a function of t i . Then ci,0 ci,1, so
c

J
j i

1 j c j,0 c j,1 .

Let t t1 , t2 , . . . , tk 1 . For j i the cubes c j,0 t and c j,1 t are independent of


ti and for j i they are independent of t i 1 . So c is a combination of degenerate
k 1-cubes and hence is degenerate.
QED
5.8. C OROLLARY. Every boundary is a cycle.
P ROOF. Suppose c b c with c degenerate. Then by Lemma 5.5 c
b c c , where we used Proposition 5.4. Lemma 5.7 says that c is
degenerate, and therefore so is c.
QED
5.9. E XAMPLE . Consider the unit circle in the plane c t ; cos 2 t, sin 2 t
with 0 t 1. This is a closed 1-cube. The circle is the boundary of the disc
of radius 1 and therefore it is reasonable to expect that c is a boundary of a 2cube. This is indeed true in the sense defined above, that c b c where c
is a constant 1-chain. (It is actually not possible to find a b such that c b;
see Exercise 5.2.) The 2-cube b is defined by shrinking c to a point, b t 1 , t2
1 t2 c t1 for t1 , t2 in the unit square. Then
b t1 , 0 o

c t1 ,
b 0, t2 L b 1, t2 L 1 t2 , 0 ,
b t1 , 1 0, 0 ,
so that b c c , where c is the constant curve located at the origin. Therefore
c b c , a boundary plus a degenerate 1-cube.
In the same way that a closed form is not necessarily exact, it may happen that
a 1-cycle is not a boundary. See Example 5.11.

5.4. STOKES THEOREM

63

5.4. Stokes theorem


In the language of chains and boundaries we can rewrite the fundamental
theorem of calculus, Theorem 4.4, as follows:

dg

g c 1 Jk g c 0 d

c 1

c 0

c 1 J

c 0

g,

i.e. c dg c g. This is the form in which the fundamental theorem of calculus


generalizes to higher dimensions. This generalization is perhaps the neatest relationship between the exterior derivative and the boundary operator. It contains
as special cases the classical integration formulas of vector calculus (Green, Gau
and Stokes) and for that reason has Stokes name attached to it, although it would
perhaps be better to call it the fundamental theorem of multivariable calculus.
5.10. T HEOREM (Stokes theorem). Let be a k 1-form on an open subset U of
Rn and let c be a k-chain in U. Then

P ROOF. By the definition of the integral and by Theorem 3.11 we have

0,1

c d


0,1

dc .

Since c is a k 1-form on 0, 1 k , it can be written as


k

gi dt1 dt2 ++|d ti ++ dtk

i 1

for certain functions g 1 , g2 , . . . , gk defined on 0, 1 k . Therefore

i 1

0,1

d gi dt1 dt2 ++|d ti ++ dtk


k

u 1 i
i 1

x
0,1

gi
dt1 dt2 ++ dtk .
ti

Changing the order of integration (cf. Remark 5.3) and subsequently applying the
fundamental theorem of calculus in one variable, formula (B.1), gives

gi
dt1 dt2 ++ dtk
k
0,1 t i

gi
dti dt1 dt2 ++|d ti ++ dtk
0,1 t i
k

0,1

k 1

gi t1 , . . . , ti 1 , 1, ti 1, . . . , tk

gi t1 , . . . , ti 1 , 0, ti 1 , . . . , tk dt1 dt2 ++|d ti ++ dtk .


The forms
gi t1 , . . . , ti

1 , 1, t i 1 , . . . , t k

dt1 dt2 ++ d ti + + dtk

gi t1 , . . . , ti

1 , 0, t i 1 , . . . , t k

dt1 dt2 ++|d ti ++ dtk

and

64

5. INTEGRATION AND STOKES THEOREM

are nothing but ci,1


, resp. ci,0
. Accordingly,

1 i

1 i

i 1 u

i 1 u
k

i 1 0,1 u

gi
dti dt1 dt2 ++d ti ++ dtk
0,1 t i
k

i 1 0,1 u

0,1

1 i

1 i

k 1


ci,1

0,1

k 1

ci,

c i,


ci,0

which proves the result.

QED

5.11. E XAMPLE . The unit circle c t


cos 2 t, sin 2 t is a 1-cycle in the

a chain in R2 it is also a boundary,


punctured plane U R2
0 . Considered
as
However, we claim that it is not a boundary in U in
as we saw in Example 5.9.
the sense that there exist no 2-chain b and no degenerate 1-chain c both contained
in U such that c b c . Indeed, suppose that c b c . Let
y dx
u other
x dy d x2 y2 be the angle form. Then c 2 by Example 4.1.On the
hand,

b c

0,

where we have used Stokes theorem, Lemma 5.5 and the fact that is closed. This
is a contradiction. The moral of this example is that the presence of the puncture
in U is responsible both for the existence of the non-exact closed 1-form (see
Example 4.6) and for the closed 1-chain c which is not a boundary. We detected
both phenomena by using Stokes theorem.
Exercises
5.1. Let U be an open subset of Rn , V an open subset of Rm and : U
map. Let c be a k-cube in U and a k-form on V. Prove that c c .
c

V a smooth

5.2. Let U be an open subset of Rn . Its boundary is a linear combination of k 1-cubes,

i ai ci .

(i) Let b be a k 1-chain in U. Its boundary is a linear combination of k-cubes,


b i ai ci . Prove that i ai 0.
(ii) Let c be a k-cube in U. Conclude that there exists no k 1-chain b in U satisfying
b c.

   

5.3. Define a 2-cube c : 0, 1 2


x1 dx3 x2 dx3 .
(i) Sketch the image of c.
(ii) Calculate both c d and
5.4. Define a 3-cube c : 0, 1
x1 dx2 dx3 . Calculate both c d and

R3 by c t1 , t2

t21 , t1 t2 , t22 , and let

and check that they are equal.

 

x1 dx2

R3 by c t1 , t2 , t3
t2 t3 , t1 t3 , t1 t2 , and let

and
check
that
they
are
equal.
c

5.5. Using polar coordinates in n dimensions (cf. Exercise 3.18) write the n 1-dimensional unit sphere S n 1 in Rn as the image of an n 1-cube c.For n 2, 3, 4, calculate
the boundary c of this cube. (The domain of c will not be the unit cube in R n 1 , but a

EXERCISES

65

rectangular block R dictated by the formula in Exercise 3.18. Choose R in such a way as to
cover the sphere as economically as possible.)
5.6. Deduce the following classical integration formulas from the generalized version
of Stokes theorem. All functions, vector fields, chains etc. are smooth and are defined in an
open subset U of Rn . (Some formulas hold only for special values of n, as indicated.)
(i)



 


 

grad g dx

g c 1


 
 
 
 


 

g c 0

for any function g and any curve c.

g
f
dx dy
x
y
and any 2-chain c. (Here n 2.)

(ii) Greens formula:


(iii) Gau formula:
any n-chain c.

(iv) Stokes formula:

div F dx1 dx2


curl F

dx

dxn

f dx

g dy for any functions f , g


dx for any vector field F and

F dx for any vector field F and any 2-chain

c. (Here n 3.)
In parts (iii) and (iv) we use the notations dx and dx explained in Section 2.5. We shall
give a geometric interpretation of the entity dx in terms of volume forms later on. (See
Corollary 8.15.)

CHAPTER 6

Manifolds
6.1. The definition
Intuitively, an n-dimensional manifold in the Euclidean space R N is a subset
that in the neighbourhood of every point looks like Rn up to smooth distortions. The formal definition is given below and is unfortunately a bit long. It will
help to consider first the basic example of the surface of the earth, which is a twodimensional sphere placed in three-dimensional space. A useful way to represent
the earth is by means of a world atlas, which is a collection of maps. Each map
depicts a portion of the world, such as a country or an ocean. The correspondence
between points on a map and points on the earths surface is not entirely faithful,
because charting a curved surface on a flat piece of paper inevitably distorts the
distances between points. But the distortions are continuous, indeed differentiable
(in most traditional cartographic projections). Maps of neighbouring areas overlap near their edges and the totality of all maps in a world atlas covers the whole
world.
An arbitrary manifold is defined similarly, as an n-dimensional world represented by an atlas consisting of maps. These maps are a special kind of
parametrizations known as embeddings.
6.1. D EFINITION . Let U be an open subset of Rn . An embedding of U into R N
is a C map : U
RN satisfying the following conditions:

     
 t );
    U, is continuous.
The image of the embedding is the set  U !  t #" t  U $ consisting
of all points of the form  t  with t  U. The inverse map  is called a chart
or coordinate map. You should think of  U  as an n-dimensional patch in R
parametrized by the map . Condition (i) means that to distinct values of the
parameter t must correspond distinct points  t  in the patch  U  . Thus the
patch  U  has no self-intersections. Condition (ii) means that for each t in U all
n columns of the Jacobi matrix D  t  must be independent. This is imposed to
prevent the occurrence of cusps and other singularities in the image  U  . Since
D  t  has N rows, this condition also implies that N % n: the target space R
must have dimension greater than or equal to that of the source space U, or else
cannot be an embedding. The column space of D  t  is called the tangent space to
the patch at the point x   t  and is denoted by T  U  ,
T  U & D  t  R  .


(i) is one-to-one (i.e. if t1


t2 , then t1
(ii) D t is one-to-one for all t U;
(iii) the inverse of , which is a map 1 : U

67

68

6. MANIFOLDS

'(

The tangent space at each point is an n-dimensional subspace of R N because D t


has n independent columns. Condition (iii) can be restated as the requirement that
if ti is any sequence of points in U such that lim i
ti exists and is equal to t
for some t U, then lim i
ti t. This is intended to avoid situations where the
image U doubles back on itself at infinity. (See Exercise 6.4 for an example.)

' ,(

)+* ' (

)+* -

'(

6.2. E XAMPLE . The picture below shows an embedding of an open rectangle


in the plane into three-space, the image of which is a portion of a torus. Try to
write a formula for such an embedding! (If we chose U too big, the image would
self-intersect and the map would not be an embedding.) For one particular value
of t the column vectors of the Jacobi matrix are also shown. As you can see, they
span the tangent plane at the image point.

./

D t e1

e3

./

./

D t e2
e2

. /

e1

e2
e1
6.3. E XAMPLE . Let U be an open subset of Rn and let f : U
map. The graph of f is the collection

Rm be a smooth

-2143 f 'tt(65877 t , U 9 .
7
Since t is an n-vector and f ' t ( an m-vector, the7 graph is a subset of R with N n : m. We claim that the graph is the image of an embedding : U 0 R . Define
t
' t ( -;3
.
f ' t ( 5
Then by definition graph f - ' U ( . Furthermore is an embedding. Indeed,
' t ( - ' t ( implies t - t , so is one-to-one. Also,
I
D ' t ( -;3
,
D f ' t (<5
graph f

6.1. THE DEFINITION

=>

69

so D t has n independent columns. Finally the inverse of is given by

? @ f =tt>6ACB
1

t,

which is continuous. Hence is an embedding.


A manifold is an object patched together out of the images of several embeddings. More precisely,
6.4. D EFINITION . An n-dimensional manifold1 (or n-manifold for short) in R N is
a subset M of R N such that for all x M there exist
an open subset V R N containing x,
an open subset U Rn ,
and an embedding : U
R N satisfying U
V M.
N
The codimension of M in R is N n. Choose t U such that t
x. Then the
tangent space to M at x is the column space of D t ,

EE

FF

= >B H
I
D= >
=>B
T M D = t >J= R > .
B
n

(Using the chain rule one can show that Tx M is independent of the choice of the
embedding .) The elements of Tx M are tangent vectors to M at x. A collection of
embeddings i : Ui
R N with Ui open in Rn and such that M is the union of all
the sets i Ui is an atlas for M.

= >

One-dimensional manifolds are called (smooth) curves, two-dimensional manifolds (smooth) surfaces, and n-manifolds in Rn 1 (smooth) hypersurfaces. In these
cases the tangent spaces are usually called tangent lines, tangent planes, and tangent
hyperplanes, respectively.
The following picture illustrates the definition. Here M is a curve in the plane,
so we have N
2 and n
1. U is an open interval in R and V is an open disc
in R2 . The map sends t to x and parametrizes the portion of the curve inside V.
Since n
1, the Jacobi matrix D t consists of a single column vector, which is
tangent to the curve at x t . The tangent line Tx M is the line spanned by this
vector.

=>
=
>
B

Tx M

U
t

x
V

Sometimes a manifold has an atlas consisting of one single chart. In that event
we can take V R N , and choose one open U Rn and an embedding : U
RN
such that M
U . However, usually one needs more than one chart to cover
a manifold. (For instance, one chart is not enough for the curve M in the picture
above.)

BB = >

1In the literature this is usually called a submanifold of Euclidean space. It is possible to define
manifolds more abstractly, without reference to a surrounding vector space. However, it turns out
that practically all abstract manifolds can be embedded into a vector space of sufficiently high dimension. Hence the abstract notion of a manifold is not substantially more general than the notion of a
submanifold of a vector space.

70

6. MANIFOLDS

6.5. E XAMPLE . An open subset U of Rn can be regarded as a manifold of dimension n (hence of codimension 0). Indeed, U is the image of the map : U
Rn given by x
x, the identity map. The tangent space to U at any point is R n
itself.

M NO

N P

NQO

6.6. E XAMPLE . Let N


n and define : Rn
R N by x1 , x2 , . . . , xn
x1 , x2 , . . . , xn , 0, 0, . . . , 0 . It is easy to check that is an embedding. Hence the
image Rn is an n-manifold in R N . (Note that Rn is just a linear subspace
isomorphic to Rn ; e.g. if N
3 and n 2 it is just the xy-plane. We shall usually
n
identify R with its image in R N .) Combining this example with the previous
one, we see that if U is any open subset of Rn , then U is a manifold in R N of
codimension N n. Its tangent space at any point is Rn .

M N

M N

M N

graph f , where f : U
Rm is a smooth map. As
6.7. E XAMPLE . Let M
shown in Example 6.3, M is the image of a single embedding : U
R n m , so
n
m
M is an n-dimensional manifold in R
, covered by a single chart. At a point
x, f x in the graph the tangent space is spanned by the columns of D. For
instance, if n m 1, M is one-dimensional and the tangent line to M at x, f x
is spanned by the vector 1, f x . This is equivalent to the well-known fact that
the slope of the tangent line to the graph at x is f x .

M M NN

O O

M TUM NN

M M NN

TVM N

graph f

Y ZVW X6[
1
f x

WX

f x

O M N\N

For n
2 and m
1, M is a surface in R3 . The tangent plane to M at a point
x, y, f x, y is spanned by the columns of D x, y , namely

M N
]a^ 1 _\` b
]^ 1 _`
0
.
M x, y N and
M
x,
y
N
0
The diagram below shows the graph of f M x, y NcO x d y R 3xy from two different
f
x

f
y

angles, together with a few points and tangent vectors. (To improve the scale the

6.1. THE DEFINITION

71

z-coordinate and the tangent vectors have been compressed by a factor of 2.)

e3

e3

e2

e1

e2
e1

f gih f
j f kg jh

R2 given by t
et cos t, sin t .
6.8. E XAMPLE . Consider the path : R
e t . Therefore
Let us check that is an embedding. Observe first that t
t
t
1
2
t1
t2 implies e
e . The exponential function is one-to-one, so t 1 t2 .
This shows that is one-to-one. The velocity vector is

f glh f g

t p sin t
mnf glh e o cos
.
cos t q sin t r
Therefore m f t gsh 0 if and only if cos t h sin t h 0, which is impossible because
cos t q sin t h 1. So m f t g+h t 0 for all t. Moreover we have t h ln e h ln j f t gkj .
Hence the inverse of is given by u f x glh ln j x j for x v f R g and so is continuous. Therefore is an embedding and f R g is a 1-manifold. The image f R g is a
spiral, which for t ewpyx converges to the origin. It winds infinitely many times
around the origin, although that is hard to see in the picture.
t

f g

f g{z}| 0 ~

Even though R is a manifold, the set R


singularity at the origin!

is not: it has a very nasty

72

6. MANIFOLDS








&

 

v  D t  D t v  D t 0  0.

6.9. E XAMPLE . An example of a manifold which cannot be covered by a single


chart is the unit sphere M S n 1 in Rn . Let U Rn 1 and let : U
Rn be the
map
1
t
2t
t 2 1 en
2
t
1
given in Exercise B.7. As we saw in that exercise, the image of is the punctured
sphere M
en , so if we let V be the open set Rn
en , then U
M V.
Also we saw that has a two-sided inverse : U
U, the stereographic projection from the north pole. Therefore is one-to-one and its inverse is continuous
(indeed, differentiable). Moreover, t
t implies D t D t v v for
all v in Rn 1 by the chain rule. Therefore, if v is in the nullspace of D t ,
Thus we see that is an embedding. To cover all of M we need a second map, for
example the inverse of the stereographic projection from the south pole. This is
also an embedding and its image is M
en
M V, where V
Rn
en .
n
This finishes the proof that M is an n 1-manifold in R .

As this example shows, the definition of a manifold can be a little awkward


to work with in practice, even for a very simple manifold. Aside from the above
examples, in practice it can be rather hard to decide whether a given subset is a
manifold using the definition alone. Fortunately there exists a more manageable
criterion for a set to be a manifold.
6.2. The regular value theorem
Definition 6.4 is based on the notion of an embedding, which can be regarded
as an explicit way of describing a manifold. However, embeddings can be hard
find in practice. Instead, manifolds are often given implicitly, by a system of m
equations in N unknowns,




1 x1 , . . . , x N
2 x1 , . . . , x N

m x1 , . . . , x N

c1 ,
c2 ,
..
.
cm .

Here the i s are smooth functions presumed to be defined on some common open
subset U of R N . Writing in the usual way

\
\
c
c
x 
, x 
, c!




... ,
x
x
c
we can abbreviate this system to a single equation
x  c.
For a fixed vector c R we denote the solution set by
c l x U x  c
x1
x2
..
.

1 x
2 x
..
.
m

1
2

6.2. THE REGULAR VALUE THEOREM

73

and call it the level set or the fibre of at c. (The notation 1 c for the solution set
is standard, but a bit unfortunate because it suggests falsely that is invertible,
which it is usually not.) If is a linear map, the system of equations is inhomogeneous linear and by linear algebra the solution set is an affine subspace of
R N . The dimension of this affine subspace is N m, provided that has rank m
(i.e. has m independent columns). We can generalize this idea to nonlinear equations as follows. We say that c
Rm is a regular value of if the Jacobi matrix
N
m
D x : R
R has rank m for all x
1 c . A vector that is not a regular
value is called a singular value. (As an extreme, though slightly silly, special case,
if 1 c is empty, then c is automatically a regular value.)
The following result is the most useful criterion for a set to be a manifold.
(Dont get carried away though, because it does not apply to every possible manifold. In other words, it is a sufficient but not a necessary criterion.) The proof uses
the following important fact from linear algebra,

nullity A

rank A

l,

valid for any k l-matrix A. Here the rank is the number of independent columns
of A (in other words the dimension of the column space A Rl ) and the nullity
is the number of independent solutions of the homogeneous equation Ax 0 (in
other words the dimension of the nullspace ker A).

6.10. T HEOREM (regular value theorem). Let U be open in R N and let : U


R be a smooth map. Suppose that c is a regular value of and that M
1 c is
N
nonempty. Then M is a manifold in R of codimension m. Its tangent space at x is the
nullspace of D x ,
m

Tx M

ker D x .

P ROOF. Let x
M. Then D x has rank m and so has m independent
columns. After relabelling the coordinates on R N we may assume the last m
columns are independent and therefore constitute an invertible m m-submatrix
A of D x . Let us put n
N m. Identify R N with Rn Rm and correspondingly write an N-vector as a pair u, v with u a n-vector and v an m-vector. Also
write x
u0 , v0 . Now refer to Appendix B.4 and observe that the submatrix
A is nothing but the partial Jacobian Dv u0 , v0 . This matrix being invertible,
by the implicit function theorem, Theorem B.4, there exist open neighbourhoods
U of u0 in Rn and V of v0 in Rm such that for each u
U there exists a unique
v
f u
V satisfying u, f u
c. The map f : U
V is C . In other words
U V
graph f is the graph of a smooth map. We conclude from ExamM
ple 6.7 that M
U V is an n-manifold, namely the image of the embedding
: U
RN given by u
u, f u . Since U V is open in R N and the above
argument is valid for every x
M, we see that M is an n-manifold. To compute
Tx M note that u
c, a constant, for all u U. Hence D u D u
0
by the chain rule. Plugging in u u0 gives

 l

 


 

D x D u l 0.
The tangent space T M is by definition the column space of D u , so every
tangent vector v to M at x is of the form v D u a for some a R . Therefore

D x v

D x D u0 a

0, i.e. Tx M

0
n

ker D x . The tangent space Tx M

74

6. MANIFOLDS

is n-dimensional (because the n columns of D u0 are independent) and so is


the nullspace of D x (because nullity D x
N m
n). Hence Tx M
ker D x .
QED

The case of one single equation (m


1) is especially important. Then D is
a single row vector and its transpose is the gradient of : D T
grad . It has
rank 1 at x if and only if it is nonzero, i.e. at least one of the partials of does
not vanish at x. The solution set of a scalar equation x
c is known as a level
hypersurface. Level hypersurfaces, especially level curves, occur frequently in all
kinds of applications. For example, isotherms in weathercharts and contour lines
in topographical maps are types of level curves.

6.11. C OROLLARY (level hypersurfaces). Let U be open in R N and let : U


R
1
be a smooth function. Suppose that M

c is nonempty and that grad x


0
for all x in M. Then M is a manifold in R N of codimension 1. Its tangent space at x is the
orthogonal complement of grad x ,

grad x\ .
6.12. E
. Let U R and x, y xy. The level curves of are
hyperbolas in the plane and the gradient is grad x y, x . The diagram
below shows a few level curves as well as the gradient vector field, which as you
Tx M

XAMPLE

can see is perpendicular to the level curves.


y

0 is the only singular value of


The gradient vanishes only at the origin, so 0
. By Corollary 6.11 this means that 1 c is a 1-manifold for c
0. The fibre
1 0 is the union of the two coordinate axes, which has a self-intersection and
so is not a manifold. However, the set 1 0
0 is a 1-manifold since the gradient is nonzero outside the origin. Think of this diagram as a topographical map
representing the surface z
x, y shown below. The level curves of are the
contour lines of the surface, obtained by intersecting the surface with horizontal
planes at different heights. As explained in Appendix B.2, the gradient points in
the direction of steepest ascent. Where the contour lines self-intersect the surface

6.2. THE REGULAR VALUE THEOREM

75

has a mountain pass or saddle point.

e3

e2
e1

6.13. E XAMPLE . A more interesting example of an equation in two variables


x3 y3 3xy
c. Here grad x
3 x 2 y, y2 x T , so grad
is x, y
T
vanishes at the origin and at 1, 1 . The corresponding values of are 0, resp.
1, which are the singular values of .
y

The level curve 1 1 is not a curve at all, but consists of the single point
1, 1 T . Here has a minimum and the surface z
x, y has a valley. The
level curve 1 0 has a self-intersection at the origin, which corresponds to a
saddle point on the surface. These features are also clearly visible in the surface
itself, which is shown in Example 6.7.

6.14. E XAMPLE . Let U R N and x


x 2 . Then grad x
2x, so as in
Example 6.12 grad vanishes only at the origin 0, which is contained in 1 0 .
So again any c 0 is a regular value of . Clearly, 1 c is empty for c 0. For
c
0, 1 c is an N 1-manifold, the sphere of radius c in R N . The tangent

76

6. MANIFOLDS

space to the sphere at x is the set of all vectors perpendicular to grad x


In other words,
Tx M x
y RN y x 0 .

2x.

Finally, 0 is a singular value (the absolute minimum) of and 1 0


0 is not
an N 1-manifold. (It happens to be a 0-manifold, though, just like the singular
fibre 1 1 in Example 6.13. So if c is a singular value, you cannot be certain
that 1 c is not a manifold. However, even if a singular fibre happens to be a
manifold, it is often of the wrong dimension.)

Here is an example of a manifold given by two equations (m

R by
x
x l;
x x

6.15. E XAMPLE . Define : R4

2
1
1 3

Then

l 2xx

D x

2).

x22
.
x2 x4

2x2
x4

0
x1

0
.
x2

If x1
0 the first and third columns of D x are independent, and if x 2
0 the
x2
0,
second and fourth columns are independent. On the other hand, if x 1
D x has rank 1 and x
0. This shows that the origin 0 in R 2 is the only
singular value of . Therefore, by the regular value theorem, for every nonzero
vector c the set 1 c is a two-manifold in R4 . For instance, M
1 10 is a
1, 0, 0, 0 T. Let us find a basis
two-manifold. Note that M contains the point x
of the tangent space Tx M. Again by the regular value theorem, this tangent space
is equal to the nullspace of

; 20

0
0

D x

0
1

0
,
0

which is equal the set of all vectors y satisfying y 1


y3
therefore given by the standard basis vectors e2 and e4 .

0. A basis of Tx M is

We now come to a more sophisticated example of a manifold determined by a


large system of equations.

6.16. E XAMPLE . Recall that an n n-matrix A is orthogonal if A T A


I. This
means that the columns (and also the rows) of A are perpendicular to one another
and have length 1. (In other words, they form an orthonormal basis of R n note
the regrettable inconsistency in the terminology.) The collection of orthogonal matrices form a group under matrix multiplication, which is usually called the orthogonal group and denoted by O n . Let us prove using the regular value theorem that
O n is a manifold. First observe that that A T A T
A T A, so A T A is a symmetric
n
n
R
matrix. In other words, if V
is the vector space of all n n-matrices and
W
C V C C T the linear subspace of all symmetric matrices, then

 A A
defines a map : V W. Clearly O n I , so to prove that O n is a
A

manifold it suffices to show that I is a regular value of . The derivative of can

EXERCISES

77

be computed by using the formula derived in Exercise B.3:

lim h1 A hB A 
lim h1 A A hA B hB A h B B A A
BA AB .
We need to show that for A O n the linear map D A : V W has rank
equal to the dimension of W. By linear algebra this amounts to showing that the
D A B

0
T

2 T

equation

BA T AB T C
(6.1)
is solvable for B, given any orthogonal A and any symmetric C. Here is a way of
1
1
guessing a solution: observe that C
C T and first try to solve BA T
2 C
2 C.
1
T
Left multiplying both sides by A and using A A
I gives B
2 CA. It is now
1
easy to check that B
CA
is
a
solution
of
equation
(6.1).
2

Exercises

U {C

R 2 by t
t sin t, 1
6.1. This is a continuation of Exercise 1.1. Define : R
T
cos t . Show that is one-to-one. Determine all t for which t
0. Prove that R is
not a manifold at these points.

6.2. Let a
0, 1 be a constant. Prove that the map : R
R 2 given by t
T
a sin t, 1 a cos t is an embedding. (This becomes easier if you first show that t
is an increasing function of t.) Graph the curve defined by .

t
a sin t

1 et
6.3. Prove that the map : R
R2 given by t
e t , et e t T is an embed2
ding. Conclude that M R is a 1-manifold. Graph the curve M. Compute the tangent
line to M at 1, 0 and try to find an equation for M.

6.4. Let I be the open interval 1,


and let : I
R 2 be the map t
3at 1
T
3
, where a is a nonzero constant. Show that is one-to-one and that
1 t
t
0 for all t
I. Is an embedding and is I a manifold? (Observe that I
is a portion of the curve studied in Exercise 1.2.)
t3 , 3at2

6.5. Define : R

R2 by

f t ,f t

f t ,f t

if t

if t

0,
0,

where f is the function given in Exercise B.6. Show that is smooth, one-to-one and that
its inverse 1 : R
R is continuous. Sketch the image of . Is R a manifold?
6.6. Define a map :

R2

R4



by
t1
t2

t31
2
t1 t2
t1 t22
t32

  

(i) Show that is one-to-one.


(ii) Show that D t is one-to-one for all t 0.
(iii) Let U be the punctured plane R2
0 . Show that : U
Conclude that U is a two-manifold in R4 .

R4 is an embedding.

78

6. MANIFOLDS




(iv) Find a basis of the tangent plane to U at the point 1, 1 .

  

 

6.7. Let : Rn
0
R be a homogeneous function of degree p as defined in Exercise B.5. Assume that is smooth and that p
0. Show that 0 is the only possible
singular value of . (Use the result of Exercise B.5.) Conclude that, if nonempty, 1 c is
an n 1-manifold for c 0.



 

 

6.8. Let x
a 1 x21 a2 x22
an x2n , where the a i are nonzero constants. Determine the regular and singular values of . For n
3 sketch the level surface 1 c for a
regular value c. (You have to distinguish between a few different cases.)
6.9. Show that the trajectories of the Lotka-Volterra system of Exercise 1.10 are onedimensional manifolds.

 

6.10. Compute the dimension of the orthogonal group O n and show that its tangent
space at the identity matrix I is the set of all antisymmetric n n-matrices.

6.11. Let V be the vector space of n


det A.
(i) Show that

n-matrices and define : V

  
n

D A B

i 1

det a1 , a2 . . . , ai

1 , bi , a i 1 , . . . , an



R by A

where a 1 , a2 , . . . , an and b1 , b2 , . . . , bn denote the column vectors of A, resp. B.


(Apply the formula of Exercise B.3 for the derivative and use the multilinearity
of the determinant.)
(ii) The special linear group is the subset of V defined by



 

SL n

V det A

 

1 .

   

" 


Show that SL n is a manifold. What is its dimension?


I, the identity matrix, we have D A B
tr B,
(iii) Show that for A
ni 1 bi,i
the trace of B. Conclude that the tangent space to SL n at I is the set of traceless
matrices, i.e. matrices A satisfying tr A 0.



(i) Let W be punctured 4-space R4

6.12.

x1 x4

x2 x3 .

0 and define : W

R by

Show that 0 is a regular value of .


(ii) Let A be a real 2 2-matrix. Show that rank A 1 if and only if det A 0 and
A 0.
(iii) Let M be the set of 2 2-matrices of rank 1. Show that M is a three-dimensional
manifold.
11 .
(iv) Compute TA M, where A
00



6.13. Define : R4

R2 by

$# %
 '&    (



 
  

x1 x2 x3 x4
x1 x2 x3 x4

   


(i) Show that D x has rank 2 unless x is of the form t 2 , t 2 , t, t 3 for some
t 0. (Compute all 2 2-subdeterminants of D and set them equal to 0.)
(ii) Show that M 1 0 is a 2-manifold (where 0 is the origin in R2 ).
(iii) Find a basis of the tangent space Tx M for all x
M with x 3
0. (The answer
depends on x.)

EXERCISES

+,
*
- + ,0/

79

) .
.

6.14. Let U be an open subset of Rn and let : U


Rm be a smooth map. Let M be
1
the manifold
c , where c is a regular value of . Let f : U
R be a smooth function.
A point x
M is called a critical point for the restricted function f M if D f x v
0 for all
tangent vectors v
Tx M. Prove that x
M is critical for f M if and only if there exist
numbers 1 , 2 , . . . , m such that
grad f x

+ ,21

1 grad 1 x

+ ,31544461

2 grad 2 x

+, /
+,

m grad m x .

(Use the characterization of Tx M given by the regular value theorem.)

6.15. Find the critical points of the function f x, y, z


given by
x2

,7/98 1 1
x

2y

3z over the circle C

1 1 /
1 /
.
/
; - =. < <>: / ?
.
+,
y2

z2

1,

0.

Where are the maxima and minima of f C?

+ ,0/ 4
+
- ,

6.16 (eigenvectors via calculus). Let A


A T be a symmetric n n-matrix and define
f : Rn
R by f x
x Ax. Let M be the unit sphere x Rn x
1 .
(i) Calculate grad f x .
(ii) Show that x M is a critical point of f M if and only if x is an eigenvector for A
of length 1.
(iii) Given an eigenvector x of length 1, show that f x is the corresponding eigenvalue of x.

CHAPTER 7

Differential forms on manifolds


7.1. First definition
There are several different ways to define differential forms on manifolds. In
this section we present a practical, workaday definition. A more theoretical approach is taken in Section 7.2.
Let M be an n-manifold in R N and let us first consider what we might mean by
a 0-form or smooth function on M. A function f : M
R is simply an assignment
of a unique number f x to each point x in M. For instance, M could be the surface
of the earth and f could represent temperature at a given time, or height above sea
level. But how would we define such a function to be differentiable? The difficulty
here is that if x is in M and e j is one of the standard basis vectors, the straight
line x he j may not be contained in M, so we cannot form the limit f x j
limh 0 f x he j
f x h.
Here is one way out of this difficulty. Because M is a manifold there exist open
R N such that the images i Ui cover M:
sets Ui in Rn and embeddings i : Ui
M
i i Ui . (Here i ranges over some unspecified, possibly infinite, index set.)
For each i we define a function f i : Ui
R by f i t
f i t , i.e. f i
i f . We
call f i the local representative of f relative to the embedding i . (For instance, if M
is the earths surface, f is temperature, and i is a map of New York State, then f i
represents a temperature chart of NY.) Since f i is defined on the open subset Ui of
Rn , it makes sense to ask whether its partial derivatives exist. We say that f is C k
if each of the local representatives f i is C k . Now suppose that x is in the overlap
of two charts. Then we have two indices i and j and vectors t
Ui and u U j
such that x i t
j u . Then we must have f x
f i t
f j u , so
1
fi t
f j u . Also i t
j u implies t i
j u and therefore f j u
f i i 1 j u . This identity must hold for all u U j such that j u
i Ui ,
1
i.e. for all u in j i Ui . We can abbreviate this by saying that

AB

F A A C BHG A BBD
E$I A B

ABE A B
E
ABE A B
ABE A B
LN M A BO
L A A BB
L A A BQB
L M

fj

A B E A A BB

A B
E J

K
K
A B E A A BB E A A BB
A BE
E L M A B
K
A BPK A B

E A L M B J f
i

on j 1 i Ui . This is a consistency condition on the functions f i imposed by the


fact that they are pullbacks of a single function f defined everywhere on M. The
map i 1 j is often called a change of coordinates and the consistency condition
is also known as the transformation law for the local representatives f i . (Pursuing
the weather chart analogy, it expresses nothing but the obvious fact that where the
maps of New York and Pennsylvania overlap, the corresponding two temperature
charts must show the same temperatures.) Conversely, the collection of all local
representatives f i determines f , because we have f x
f i i 1 x if x i Ui .

A B E A L A BB K A B

81

82

7. DIFFERENTIAL FORMS ON MANIFOLDS

(That is to say, if we have a complete set of weather charts for the whole world, we
know the temperature everywhere.)
Following this cue we formulate the following definition.
7.1. D EFINITION . A differential form of degree k, or simply a k-form, on M is a
collection of k-forms i on Ui satisfying the transformation law

U S S WW
Y S W

RTS U V WX
i

(7.1)

on j 1 i Ui . We call i the local representative of relative to the embedding


i and denote it by i
i . The collection of all k-forms on M is denoted by
k M .

R X

This definition is rather indirect, but it works really well if a specific atlas for
the manifold M is known. Definition 7.1 is particularly tractible if M is the image
of a single embedding : U
R N . In that case the compatibility relation (7.1) is
vacuous and a k-form on M is determined by one single representative, a k-form
on U.
Sometimes it is useful to write the transformation law (7.1) in components. We
can do this by appealing to Theorem 3.12. If

f I dt I

and

are two local representatives for , then


gJ

U S S WW

R S U V W X f
1

g J dt J
J

S U V W

det D i

j I,J .

on j 1 i Ui .
Just like forms on Rn , forms on a manifold can be added, multiplied, differentiated and integrated. For example, suppose is a k-form and an l-form on M.
Suppose i , resp. i , is the local representative of , resp. , relative to an embedding i : Ui
M. Then we define the product by setting i ii . To see
that this definition makes sense, we check that the forms i satisfy the transformation law (7.1):

R[S U V W X S U V W X R\S U V W X S W]RTS U V W X .


Here we have used the multiplicative property of pullbacks, Proposition 3.10(ii).
Similarly, the exterior derivative of is defined by setting S d W R d . As before,
let us check that the forms S d W satisfy the transformation law (7.1):
S d W R d R d S U V W X RTS U V W X d RTS U V W X S d W ,
j

j j

i i

where we used Theorem 3.11.

7.2. Second definition


This section presents some of the algebraic underpinnings of the theory of
differential forms. This branch of algebra, now called exterior or alternating algebra
was invented by Gramann in the mid-nineteenth century and is a prerequisite
for much of the more advanced literature on the subject.

7.2. SECOND DEFINITION

83

Covectors. Before giving a rigorous definition of differential forms on manifolds we need to be more precise about the definition of a differential form on R n .
Recall that Rn is the collection of all column vectors

_```a x bQccc
x
x^
.. .
. d
1
2

xn

Let U be an open subset of Rn . The definition of a 0-form on U requires no further


clarification: it is simply a function on U. Formally, a 1-form on U can be defined
as a row vector

f1, f2, . . . , fn

^Te

e fg

whose entries are functions on U. The form is called constant if the entries f 1 , . . . ,
f n are constant. The set of constant row vectors is denoted by R n and is called
the dual of Rn . Constant 1-forms are also known as covariant vectors or covectors
and arbitrary 1-forms as covariant vector fields or covector fields. By definition dx i is
the constant 1-form
0, . . . , 0, 1, 0, . . . , 0 ,
dxi eiT

^Te

the transpose of ei , the i-th standard basis vector of


written as

Rn .

Every 1-form can thus be

^\e f , f , . . . , f f ^ h f dx .
n

i 1

Using this formalism we can write for any smooth function g on U


dg

^ h xg dx ^i xg , xg , . . . , xg j ,
n

i 1

so dg is simply the Jacobi matrix Dg of g! (This is the reason that many authors
use the notation dg for the Jacobi matrix.)
We would like to extend the notions of covectors and 1-forms to vector spaces
other than Rn . To see how, let us start by observing that a row vector y is nothing
but a 1 n-matrix. We can multiply it by a column vector x to obtain a number,

_```a x b ccc
x
yx ^Te y , y , . . . , y f
^ h y x .
..
. d
x
Obviously we have y e c x l c x f ^ c yx l c yx . Thus a row vector can be
viewed as a linear map which sends column vectors in R to one-dimensional
vectors (scalars) in R ^ R.
This motivates the following definition. If V is any vector space over the real
numbers (for example R or a subspace of R ), then V g , the dual of V, is the set
of linear maps from V to R. Elements of V g are called dual vectors or covectors or
linear functionals. The dual is a vector space in its own right: if and are in V g
we define l and c by setting e l f e v f ^ e v f l e v f for all v m V
and e c f e v f ^ c e v f .
1

i i

i 1

1 1

2 2

84

7. DIFFERENTIAL FORMS ON MANIFOLDS

oqp r s
p r
o stn
u os
o v s n o s v o s
7.3. E
. Let V n R and fix v w V. Define o x sxn v y x, where y is the
standard inner product on R . Then is a linear functional on V.
n

7.2. E XAMPLE . Let V


C 0 a, b , R , the collection of all continuous realvalued functions on a closed and bounded interval a, b . A linear combination
of continuous functions is continuous, so V is a vector space. Define f
b
c2 f 2
c1 f 1
c2 f 2 , so is a linear functional
a f x dx. Then c 1 f 1
on V.
n

XAMPLE

Now suppose that V is a vector space of finite dimension n and choose a basis
v1 , v2 , . . . , vn of V. Then every vector v
V can be written in a unique way as
a linear combination j c j v j . Define a covector i
V by i v
ci . In other
words, i is determined by the rule

o s]n

i v j

n}| 10

i, j

if i
if i

We call i the i-th coordinate function.

zn n

o s{n

w z
j,
j.

n~

7.4. L EMMA . The coordinate functions 1 , 2 , . . . , n constitute a basis of V . Hence


dim V
n dim V.

w z

P ROOF. Let V . We need to write as a linear combination ni 1 ci i .


Assuming for the moment that this is possible, we can apply both sides to the
vector v j to obtain

o xs n c o v s]n c n c .
(7.2)

So c n o v s is the only possible choice for the coefficient c . To show that this
choice of coefficients works, let us define n o v s . Then by equation
(7.2), o v s]n o v s for all j, so n , i.e. n o v s . We have proved that
QED
every w V z can be written uniquely as a linear combination of the .
The basis , , . . . , of V z is said to be dual to the basis v , v , . . . , v
of V.
7.5. E
. Consider R with standard basis e , . . . , e . Then dx o e sn
e e n , so the dual basis of o R sz is dx , dx , . . . , dx .
Dual bases come in handy when writing the matrix of a linear map. Let
L : V W be a linear map between abstract vector spaces V and W. To write
the matrix of L we need to start by picking a basis v , v , . . . , v of V and a basis
w , w , . . . , w of W. Then for each j n 1, 2, . . . , n the vector Lv can be expanded
uniquely in terms of the ws: Lv n l w . The m n numbers l make up
the matrix of L relative to the two bases of V and W.
7.6. L
. Let , , . . . , w V z be the dual basis of v , v , . . . , v and ,
, . . . , w W z the dual basis of w , w , . . . , w . Then the i, j-th matrix element of a
linear map L : V W is equal to l n o Lv s .
P
. We have Lv n l w , so
o Lv s]n l o w sn l n l ,

o
that is l n Lv s .
QED
n

vj

i i

i 1

i 1

i i, j

n
i 1

n
i 1

XAMPLE

T
i j

i, j

EMMA

ROOF

i, j

i, j
m
k 1 k, j
m

m
i 1 i, j

i, j

k 1

k, j i

k 1

k, j ik

i, j

7.2. SECOND DEFINITION

85

Multilinear algebra. Let V be a vector space and let V k denote the Cartesian product V
V (k times). Thus an element of V k is an ordered k-tuple
v1 , v2 , . . . , vk of vectors in V. A k-multilinear function on V is a function : V k
R which is linear in each vector, i.e.

c v , . . . , v

c v , v , . . . , v c v , v , . . . , v , . . . , v
for all scalars c, c and all vectors
v , v , . . . , v , v , . . . , v .
R and let x, y
x y, the inner product of x and
7.7. E
. Let V
y. Then is bilinear (i.e. 2-multilinear).

vw vw
7.8. E
. Let V
R ,k
2. The function v, w
v w v w is bilinear on R .

7.9. E
. Let V
R ,k
n. The determinant det v v , , . . . , v is an
v1 , v2 , . . . , cvi

XAMPLE

XAMPLE

XAMPLE

n-multilinear function on

1 2

Rn .

A k-multilinear function is alternating or antisymmetric if it has the alternating


property,

v , . . . , v , . . . , v , . . . , v .
More generally, if is alternating, then for any permutation S we have


v ,...,v
sign v , . . . , v .
of Example 7.7 is bilinear, but it is not al7.10. E
. The inner product
ternating. Indeed it is symmetric: y x x y. The bilinear function of Example 7.8
is alternating, and so is the determinant function of Example 7.9.
v1 , . . . , v j , . . . , v i , . . . , v k

XAMPLE

Here is a useful trick to generate alternating k-multilinear functions starting


from k covectors 1 , 2 , . . . , k V . The (wedge) product is the function

: V R
defined by

v , v , . . . , v
det v  .
(The determinant on the right is a k k-determinant.) It follows from the multilinearity and the alternating property of the determinant that is an alternating k-multilinear function. The wedge product is often denoted by
to distinguish it from other products, such as the tensor product defined
in Exercise 7.6.

The collection
of all alternating k-multilinear functions is denoted by A V.
For k
1 the alternating property is vacuous, so an alternating 1-multilinear
function is nothing but a linear function. Thus A V V .
k 0 a k-multilinear function is defined to be a single number. Thus
For
1 2

1 2

1 i, j k

1 2

A0 V R.
For any k, k-multilinear functions can be added and scalar-multiplied just like
ordinary linear functions, so the set A k V forms a vector space.
There is a nice way to construct a basis of the vector space A k V starting from
a basis v1 , . . . , vn of V. The idea is to take wedge products of dual basis vectors.

86

7. DIFFERENTIAL FORMS ON MANIFOLDS

i , i , . . . , i be

A V,
v T v , v , . . . , v V .
7.11. E
. Let V R with standard basis e , e , e . The dual basis
of R is dx , dx , dx . Let k 2 and I [ 1, 2 , J T 2, 3 . Then
dx e }
dxdx ee dxdx ee 10 01 1,
dx e
dxdx ee dxdx ee 01 00 0,
dx e }
dxdx ee dxdx ee 00 10 0,
dx e }
dxdx ee dxdx ee 10 01 1.

Let 1 , . . . , n be the corresponding dual basis of V . Let I


an increasing multi-index, i.e. 1 i 1 i2
ik n. Write
I

i1 i2

i1

ik

i2

ik

XAMPLE

This example generalizes as follows.

7.12. L EMMA . Let I and J be increasing multi-indices of degree k. Then

} 10 ifif II J,J.
P
. Let I T i , . . . , i and J T j , . . . , j . Then
... . . . ...
. . . 1 if I J.
v det v 
..
...
. . . .
If I J, then i j for some l. Choose l as small as possible, so that i j for
m l. There are two cases: i j and i j . If i j , then i j j
7 j because J is increasing, so all entries in the determinant with m0 forl
are 0. For m l we have j i i because I is increasing, so
m l. In other words the l-th row in the determinant is 0 and hence v 0.
If i j we find that the l-th column in the determinant is 0 and therefore again
v 0.
QED
I vJ

ROOF

ir

I,J

js

1 r,s k

i1 , j1

i1 , jk

il , j1

il , jk

ik , j1

ik , jk

il , jm

il , jm

l 1

We need one further technical result before showing that the functions I are
a basis of Ak V.

7.13. L EMMA . Let


of degree k. Then 0.

A k V. Suppose v I

P ROOF. The assumption implies

vi 1 , . . . , v i k

0 for all increasing multi-indices I

(7.3)

7.2. SECOND DEFINITION

87

for all multi-indices i1 , . . . , ik , because of the alternating property. We need to


show that w1 , . . . , wk
0 for arbitrary vectors w1 , . . . , wk . We can expand the
wi using the basis:

w1
wk
Therefore by multilinearity

w1 , . . . , w k

a11 v1

a1k vk ,

ak1 v1

akk vk .

..
.

x a a v , . . . , v .
k

i1 1

ik 1

1i 1

i1

ki k

ik

Each term in the right-hand side is 0 by equation (7.3).

QED

7.14. T HEOREM . Let V be an n-dimensional vector space with basis v 1 , . . . , vn .


Let 1 , . . . , n be the corresponding dual basis of V . Then the alternating k-multilinear
functions I
i1
ik , where I ranges over the set of all increasing multi-indices of
n
degree k, form a basis of A k V. Hence dim Ak V
k .


P
. The proof is closely analogous to that of Lemma 7.4. Let A V.
We need to write as a linear combination c . Assuming for the moment
that this is possible, we can apply both sides to the k-tuple of vectors v . Using
k

ROOF

I I I

Lemma 7.12 we obtain

vJ

cI I

v ]
J

c I I,J

cJ.

>

> {

7.15. E
. Let V R with standard basis e , . . . , e . The dual basis
of R is dx , . . . , dx . Therefore A V has a basis consisting of all k-multilinear
functions of the form
dx dx dx dx ,
with 1 i \H i n. Hence a general alternating k-multilinear function
on R looks like
a dx ,
with a constant. By Lemma 7.12, e a dx e a a , so a is
equal to e .
An arbitary k-form on a region U in R is now defined as a choice of an
alternating k-multilinear function for each x U; hence it looks like
f x dx , where the coefficients f are functions on U. We shall abbreviate this
I

So c J
v J is the only possible choice for the coefficient c J . To show that this
choice of coefficients works, let us define
I v I I . Then for all increasing
multi-indices I we have v I
vI
vI
vI
0. Applying Lemma
7.13 to we find
0. In other words, I v I I . We have proved
that every V can be written uniquely as a linear combination of the i . QED
n

XAMPLE

i1

i2

ik

I I,J

I I

to

f I dx I ,
I

88

7. DIFFERENTIAL FORMS ON MANIFOLDS

and we shall always assume the coefficients f I to be smooth functions. By Example


7.15 we can express the coefficients as f I
e I (which is to be interpreted as
fI x
x e I for all x).

Pullbacks re-examined. In the light of this new definition we can give a fresh
interpretation of a pullback. This will be useful in our study of forms on manifolds.
Let U and V be open subsets of Rn , resp. Rm , and : U
V a smooth map. For a
k V define the pullback
k U by
k-form



v , v , . . . , v ] D x v , D x v , . . . , D x v .
Let us check that this formula agrees with the old definition. Suppose f dy
and g dx . What is the relationship between g and f ? We use g
e , our new definition of pullback and the definition of the wedge product
to get
g x ] e x D x e , D x e , . . . , D x e
f x dy D x e , D x e , . . . , D x e
f x det dy D x e  .
By Lemma 7.6 the number dy D x e is the i j -matrix entry of the Jacobi matrix D x (with respect to the standard basis e , e , . . . , e of R and the standard
basis e , e , . . . , e of R ). In other words, g x f x det D x . This
x

I I

j1

j1

ir

jk

1 r,s k

r s

js

jk

j2

js

ir

j2

I,J

formula is identical to the one in Theorem 3.12 and therefore our new definition
agrees with the old!

Forms on manifolds. Let M be an n-dimensional manifold in R N . For each


point x in M the tangent space Tx M is an n-dimensional linear subspace of R N .
A differential form of degree k or a k-form on M is a choice of an alternating kmultilinear map x on the vector space Tx M, one for each x M. This alternating
map x is required to depend smoothly on x in the following sense. According to
the definition of a manifold, for each x M there exists an embedding : U
RN
N
such that U
M V for some open set V in R containing x. The tangent
space at x is then Tx M
D t Rn , where t U is chosen such that t
x.
The pullback of under the local parametrization is defined by


v , v , . . . , v D t v , D t v , . . . , D t v .
Then is a k-form on U, an open subset of R , so f dt for certain
functions f defined on U. We will require the functions f to be smooth. (The
form f dt is the local representative of relative to the embedding , as
introduced in Section 7.1.) To recapitulate:
7.16. D
. A k-form on M is a choice, for each x M, of an alternating k-multilinear map on T M, which depends smoothly on x.
t

I I

EFINITION

The book [BT82] describes a k-form as an animal that inhabits a world M,


eats ordered k-tuples of tangent vectors, and spits out numbers.
7.17. E XAMPLE . Let M be a one-dimensional manifold in R N . Let us choose an
orientation (direction) on M. A tangent vector to M is positive if it points in the

EXERCISES

89

same direction as the orientation and negative if it points in the opposite direction.
Define a 1-form on M as follows. For x M and a tangent vector v Tx M put

vv

if v is positive,
if v is negative.

x v

This form is the element of arc length of M. We shall see in Chapter 8 how to generalize it to higher-dimensional manifolds and in Chapter 9 how to use it to calculate
arc lengths and volumes.

We can calculate the local representative of a k-form for any embedding


: U
R N parametrizing a portion of M. Suppose we had two different such
M and j : U j
M, such that x is contained in both Wi
embeddings i : Ui
i Ui and W j
j U j . How do the local expressions i
i and j
j for compare? To answer this question, consider the coordinate change map

, which maps W W to W W . From and


we recover the transformation law (7.1)
.
j

This shows that Definitions 7.1 and 7.16 of differential forms on a manifold are
equivalent.

7.1. The vectors e 1

e2 and e1

Exercises

e2 form a basis of R2 . What is the dual basis of R2 ?

7.2. Let v1 , v2 , . . . , vn be a basis of Rn and let 1 , 2 , . . . , n be the dual basis of


Rn . Let A be an invertible n n-matrix. Then by elementary linear algebra the set of
vectors Av1 , . . . , Avn is also a basis of Rn . Show that the corresponding dual basis is the
set of row vectors 1 A 1 , 2 A 1 , . . . , n A 1 .

7.3. Suppose that is a bilinear function on a vector space V satisfying v, v


0 for
all vectors v V. Prove that is alternating. Generalize this observation to k-multilinear
functions.
7.4. Show that the bilinear function of Example 7.8 is equal to dx 1 dx2

dx3 dx4 .

7.5. The wedge product is a generalization of the cross product to arbitrary dimensions
in the sense that
T
x y
xT yT

for all x, y R3 . Prove this formula. (Interpretation: x and y are column vectors, x T and yT
are row vectors, x T yT is a 2-form on R3 , xT yT is a 1-form, i.e. a row vector. So both
sides of the formula represent column vectors.)
7.6. Let V be a vector space and let 1 , 2 , . . . , k
product is the function

1
defined by

5 
Q6
1

Show that 1

Q6
2

k v1 , v2 , . . . , vk

k : V k

V be covectors. Their tensor

0  
1 v1 2 v 2

k is a k-multilinear function.

k vk .

90

by

7.7. Let : V k

7. DIFFERENTIAL FORMS ON MANIFOLDS

R be a k-multilinear function. Define a new function Alt : V k

Alt v1 , v2 , . . . , vk 

1
sign v  1  , v  2  , . . . , v  k  .
k!
S
k

Prove the following.


(i) Alt is an alternating k-multilinear function.
(ii) Alt  if is alternating.
(iii) Alt Alt  Alt for all k-multilinear .
(iv) Let 1 , 2 , . . . , k  V  . Then
1
Alt 1
2
  
k .
1 2    k 
k!

7.8. Show that det v1 , v2 , . . . , vn  dx1 dx2    dxn v1 , v2 , . . . , vn for all vectors v1 ,
v2 , . . . , vn  Rn . In short,
det  dx1 dx2   dxn .

7.9. Let V and W be vector spaces and L : V


L  L  for all covectors ,  W  .

W a linear map. Show that L  

CHAPTER 8

Volume forms
8.1. n-Dimensional volume in R N
Let a1 , a2 , . . . , an be vectors in R N . The block or parallelepiped spanned by these
vectors is the set of all vectors of the form ni 1 ci ai , where the coefficients c i range
over the unit interval  0, 1  . For n  1 this is also called a line segment and for n  2
a parallelogram. We will need a formula for the volume of a block. If n  N there
is no coherent way of defining an orientation on all n-blocks in R N , so this volume
will be not an oriented but an absolute volume. We approach this problem in a
similar way as the problem of defining the determinant, namely by imposing a
few reasonable axioms.
tion

8.1. D EFINITION . An absolute n-dimensional Euclidean volume function is a funcvol n : R N

R N 
  RN

n times

with the following properties:


(i) homogeneity:

! c

voln  a1 , a2 , . . . , cai , . . . , an 

voln  a1 , a2 , . . . , an 

for all scalars c and all vectors a1 , a2 , . . . , an ;


(ii) invariance under shear transformations:
vol n  a1 , . . . , ai

"

ca j , . . . , a j , . . . , an 

voln  a1 , . . . , a j , . . . , ai , . . . , an 

for all scalars c and any i  # j;


(iii) invariance under Euclidean motions:
voln  Qa1 , . . . , Qan 

for all orthogonal matrices Q;


(iv) normalization: vol n  e1 , e2 , . . . , en 

voln  a1 , . . . , an 

1.

We shall shortly see that these axioms uniquely determine the n-dimensional
volume function.
8.2. L EMMA .
(i) vol n  a1 , a2 , . . . , an  %$ a1 $&$ a2 $  $ an $ if a1 , a2 , . . . ,
an are orthogonal vectors.
(ii) vol n  a1 , a2 , . . . , an   0 if the vectors a1 , a2 , . . . , an are dependent.
P ROOF. Suppose a1 , a2 , . . . , an are orthogonal. First assume they are nonzero.
Then we can define qi '$ ai $( 1 ai . The vectors q1 , q2 , . . . , qn are orthonormal.
Complete them to an orthonormal basis q1 , q2 . . . , qn , qn ) 1 , . . . , qN of R N . Let
91

92

8. VOLUME FORMS

Q be the matrix whose i-th column is qi . Then Q is orthogonal and Qei


Therefore
vol n + a1 , a2 , . . . , an ,

*.* .
*.*.-

qi .

a1 -&- a2 -0///1- an - voln + q1 , q2 , . . . , qn ,

by Axiom (i)

a1 -&- a2 -0///1- an - voln + e1 , e2 , . . . , en ,

by Axiom (iii)

a1 -&- a2 -0///1- an - voln + Qe1 , Qe2 , . . . , Qen ,


a1 -&- a2 -0///1- an -

by Axiom (iv),

which proves part (i) if all ai are nonzero. If one of the ai is 0, the vectors a1 , a2 , . . . ,
an are dependent, so the statement follows from part (ii), which we prove next.
Assume a1 , a2 , . . . , an are dependent. For simplicity suppose a1 is a linear
combination of the other vectors, a1 * ni2 2 ci ai . By repeatedly applying Axiom
(ii) we get
vol n + a1 , a2 , . . . , an ,

vol n

c i ai , a 2 , . . . , a n 4
23

i 2

vol n

ci ai , a2 , . . . , an 4 *5///6*
2

i 3

vol n + 0, a2 , . . . , an , .

Now by Axiom (i),


vol n + 0, a2 , . . . , an ,

voln + 0 0, a2 , . . . , an ,

0 voln + 0, a2 , . . . , an ,

which proves property (ii).

0,
QED

This brings us to the volume formula. We can form a matrix A out of the
column vectors a1 , a2 , . . . , an . It does not make sense to take det A because A is
not square, unless n * N. However, the product A T A is square and we can take
its determinant.
8.3. T HEOREM . There exists a unique n-dimensional volume function on R N . Let
a1 , a2 , . . . , an 7 R N and let A be the N 8 n-matrix whose i-th column is a i . Then
voln + a1 , a2 , . . . , an ,

*59

det + A T A , .

P ROOF. We leave it to the reader to check that the function : det + A T A , satisfies the axioms for a n-dimensional volume function on R N . (See Exercise 8.2.)
Here we prove only the uniqueness part of the theorem.
Case 1. First assume that a1 , a2 , . . . , an are orthogonal. Then A T A is a diagonal
matrix. Its i-th diagonal entry is - ai - 2 , so : det + A T A , *.- a1 -&- a2 -0///;- an - , which
is equal to voln + a1 , a2 , . . . , an , by Lemma 8.2(i).
Case 2. Next assume that a1 , a2 , . . . , an are dependent. Then the matrix A
has a nontrivial nullspace, i.e. there exists a nonzero n-vector v such that Av *
0. But then A T Av * 0, so the columns of A T A are dependent as well. Since
A T A is square, this implies det A T A * 0, so : det + A T A , * 0, which is equal to
voln + a1 , a2 , . . . , an , by Lemma 8.2(ii).
Case 3. Finally consider an arbitrary sequence of independent vectors a1 ,
a2 , . . . , an . This sequence can be transformed into an orthogonal sequence v 1 ,
v2 , . . . , vn by the Gram-Schmidt process. This works as follows: let b 1 * 0 and for
i < 1 let bi be the orthogonal projection of ai onto the span of a1 , a2 , . . . , ai = 1 ; then

8.1. n-DIMENSIONAL VOLUME IN RN

vi > ai ? bi . (See illustration below.) Let V be the N


is vi . Then by repeated applications of Axiom (ii),
vol n A a1 , a2 , . . . , an B

>

vol n A v1 , a2 , . . . , an B

>

>

93

n-matrix whose i-th column

vol n A v1 , v2 , . . . , an B

voln A v1 , v2 , . . . , vn B

>5D

>5CCC

det A V T V B , (8.1)

where the last equality follows from Case 1. Since vi > ai ? bi , where bi is a linear
combination of a1 , a2 , . . . , ai E 1, we have V > AU, where U is a n @ n-matrix of the
form
FGG
JK
I
CCC I KK
GG 1 I
CCC I KK
GH 0 1 I
0 0 1
CCC I .
U>
.. ..
..
. I L
. .
0 0 CCC
0 1
Note that U has determinant 1. This implies that V T V
det A A T A B

>

det U T det A A T A B det U

>

U T A T AU and

det A U T A T AU B

>

Using formula (8.1) we get vol n A a1 , a2 , . . . , an B

>NM

>

det A V T V B .

det A A T A B .

QED

The Gram-Schmidt process transforms a sequence of n independent vectors


a1 , a2 , . . . , an into an orthogonal sequence v 1 , v2 , . . . , vn . (The horizontal floor
represents the plane spanned by a1 and a2 .) The block spanned by the as has the
same volume as the rectangular block spanned by the vs.
a3
v3

a3
a2

a2
a1

v2

b2

a1

b3

a3

v1

v3

a2
a1

v2

v1
For n

>

N Theorem 8.3 gives the following result.

8.4. C OROLLARY. Let a1 , a2 , . . . , an be vectors in Rn and let A be the n


whose i-th column is ai . Then voln A a1 , a2 , . . . , an B >.P det A P .

n-matrix

94

8. VOLUME FORMS

P ROOF. A is square, so det Q A T A RTS det A T det A SUQ det A R 2 by Theorem


3.7(ii) and therefore vol n Q a1 , a2 , . . . , an RVSXW Q det A R 2 SZY det A Y by Theorem 8.3.
QED
8.2. Orientations
Oriented vector spaces. You are probably familiar with orientations on vector spaces of dimension [ 3. An orientation of a line is an assignment of a direction. An orientation of a plane is a choice of a direction of rotation, clockwise
versus counterclockwise. An orientation of a three-dimensional space is a choice
of handedness, i.e. a choice of a right-hand rule versus a left-hand rule.
These notions can be generalized as follows. Let V be an n-dimensional vector space over the real numbers. Suppose that \]S^Q v1 , v2 , . . . , vn R and \`_aS
Q v1_ , v2_ , . . . , vn_ R are two ordered bases of V. Then we can write vi_ S j ai, jv j and
vi S j bi, j v _ j for suitable coefficients a i, j and bi, j. The n b n-matrices A SZQ a i, j R
and B ScQ bi, j R satisfy AB S BA S I and are therefore invertible. We say that the
bases \ and \ _ define the same orientation of V if det A d 0. If det A e 0, the two
bases define opposite orientations.
For instance, if Q v1_ , v2_ , . . . , vn_ R&S5Q v2 , v1 , . . . , vn R , then

fgg
gg

gh

0
1
0
..
.

1
0
0
..
.

0
0
1
..
.

...
...
...
..
.
...

ijj

0
0
0
..
.k

jj
,

so det A Sml 1. Hence the ordered bases Q v2 , v1 , . . . , vn R and Q v1 , v2 , . . . , vn R define


opposite orientations.
We know now what it means for two bases to have the same orientation, but
how do we define the concept of an orientation itself? In typical mathematicians
fashion we define the orientation of V determined by the basis \ to be the collection of all ordered bases that have the same orientation as \ . (There is an analogous definition of the number 1, namely as the collection of all sets that contain
one
The orientation determined by \^SnQ v1 , v2 , . . . , vn R is denoted
o element.)
o
o
o by
\qp or v1 , v2 , . . . , vn p . So if \ and \ _ define theo same orientation
q
\
r
p
S
\ _p.
then
o
If they define opposite orientations we write \qpsScl \ _ p . Because the determinant of an invertible matrix is either positive or negative, there are two possible
orientations of V. An oriented vector space is a vector space together with a choice
of an orientation. This preferred orientation is then called positive.
For n S 0 we need to make a special definition, because a zero-dimensional
space has an empty basis. In this case we define an orientation of V to be a choice
of sign, t or l .

8.5. E XAMPLE . The standard orientation on Rn is the orientation e1 , . . . , en p defined by the standard ordered basis Q e1 , . . . , en R . We shall always use this orientation on Rn .
Maps and orientations. Let V and W be oriented vector spaces of the same
dimension and let L : V u W be an invertible linear map. Choose a positively
oriented basis Q v1 , v2 , . . . , vn R of V. Because L is invertible, the ordered n-tuple

8.2. ORIENTATIONS

95

Lv1 , Lv2 , . . . , Lvn w is an ordered basis of W. If this basis is positively, resp. negatively, oriented we say that L is orientation-preserving, resp. orientation-reversing.
v
This definition does not depend on the choice of the basis, for if v v1x , v2x , . . . , vnx w is
another positively oriented basis of V, then vix y j ai, j v j with det ai, j w{z 0. Therev
fore Lvix y L | j ai, jv j } y j ai, j Lv j , and hence the two bases Lv 1 , Lv2 , . . . , Lvn w
v
and Lv1x , Lv2x , . . . , Lvnx w of W determine the same orientation.
Oriented manifolds. Now let M be a manifold. We define an orientation of
M to be a choice of an orientation for each tangent space Tx M which varies continuously over M. Continuous means that for every x ~ M there
v exists a loM, with W open in Rn and x ~ W w , such that
cal parametrization : W 
Dy : Rn  Ty M preserves the orientation for all y ~ W. (Here Rn is equipped
with its standard orientation.) A manifold is orientable if it possesses an orientation; it is oriented if a specific orientation has been chosen.
Hypersurfaces. The case of a hypersurface, a manifold of codimension 1, is
particularly instructive. A unit normal
vector field on av manifold M in R n is a smooth
v
n
function n : M  R such that n x w Tx M and n x w y 1 for all x ~ M.

8.6. P ROPOSITION . A hypersurface in Rn is orientable if and only if it possesses a


unit normal vector field.

n
P ROOF. Let M
v be a hypersurface in R . Suppose M possesses a unit normal
vectorv field.
Let v1 , v2 , . . . , vn 1 w be an ordered basis
v
v of Tx M for some x ~ M.
n , because n x
n
x
,
v
,
v
,
.
.
.
,
v
is
a
basis
of
R
Then
w
w
w& vi for all i. We say that
1
2
n

1
v
v v
v1 , v2 , . . . , vn 1 w is positively oriented if n x w , v1 , v2 , . . . , vn 1 w is a positively oriented basis of Rn . This defines an orientation on M, called the orientation induced
by the normal vector field n.
Conversely, let us suppose that M is an oriented hypersurface in R n . For each
x ~ M the tangent space Tx M is n 1-dimensional, so its orthogonal complement
v
Tx M w is a line. There are therefore precisely two vectors of length 1 which are
perpendicular
to Tx M. We can pick a preferred unit normal vector as follows. Let
v
v1 , v2 , . . . , vn 1 w be a positively oriented
basis of Tv x M.
v
v The positive unit normal
vector is that unit normal vector n x w that makes n x w , v1 , v2 , . . . , vn 1 w a posiv
tively oriented basis of Rn . In Exercise 8.8 you will be asked to check that n x w
depends smoothly on x. In this way we have produced a unit normal vector field
on M.
QED

8.7. E XAMPLE . Let us regard Rn 1 as the subspace of Rn spanned by the first


n 1 standard basis vectors e1 , e2 , . . . , en 1 . The standard
orientation on Rn is

1
e1 , e2 , . . . , en , and the standard orientation on R
is e1 , e2 , . . . , en 1 . Since

e n , e1 , e2 , . . . , e n 1
v
by Exercise 8.5, the positive unit normal to Rn 1 in Rn is 1 w n
e1 , e 2 , . . . , e n

1w

n 1

1e

M in Rn

n.

The positive unit normal on an oriented hypersurface


can be regarded
as a map n from M into the unit sphere S n 1 , which is often called the Gau map of
M. The unit normal enables one to distinguish between two sides of M: the direction of n is out or up; the opposite direction is in or down. For this reason
orientable hypersurfaces are often called two-sided, whereas the nonorientable ones
are called one-sided. Let us show that a hypersurface given by a single equation is
always orientable.

96

8. VOLUME FORMS

8.8. P ROPOSITION . Let U be open in Rn and let : U R be a smooth function.


Let c be a regular value of . Then the manifold 1 c has a unit normal vector field
given by n x & grad x grad x ; and is therefore orientable.

1 c is a hypersurface
in Rn (if nonempty), and also that Tx M ker Dx c grad x  . The function
n x { grad x grad x ; therefore defines a unit normal vector field on M.
Appealing to Proposition 8.6 we conclude that M is orientable.
QED
P ROOF. The regular value theorem tells us that M

8.9. E XAMPLE . Taking x % x 2 and c r2 we obtain that the n


sphere of radius r about the origin is orientable. The unit normal is
n x

grad x grad x ;

1 -

x x .

8.3. Volume forms


Now let M be an oriented n-manifold in R N . Choose a collection of embeddings i : Ui R N with Ui open in Rn such that M i i Ui and such that
Di t : Rn Tx M is orientation-preserving for all t Ui . The volume form M ,
also denoted by , is the n-form on M whose local representative relative to the
embedding i is defined by

det Di t T Di t dt1 dt2



dtn .

By Theorem 8.3 the square-root factor measures the volume of the n-dimensional
block in the tangent space Tx M spanned by the columns of D i t , the Jacobi matrix of i at t. Hence you should think of as measuring the volume of infinitesimal blocks inside M.
8.10. T HEOREM . For any oriented n-manifold M in R N the volume form M is a
well-defined n-form.
P ROOF. To show that is well-defined we need to check that its local representatives satisfy the transformation law (7.1). So let us put i 1 j and substitute t u into i . Since each of the embeddings i is orientation-preserving,
we have det D 0. Hence by Theorem 3.13 we have

dt1 dt2  dtn &

det D u du1 du2



dun

! det D u ; du1 du2 

dun .

Therefore

det D u; du1 du2  dun


det D u T det Di u T Di u  det D u du1 du2 
det Di u D u T Di u D u du1 du2  dun
det D j u T D j u du1 du2  dun j ,

det Di u T Di u

where in the second to last identity we applied the chain rule.

dun

QED

For n 1 the volume form is usually called the element of arc length, for n 2,
the element of surface area, and for n 3, the volume element. Traditionally these are
denoted by ds, dA, and dV, respectively. Dont be misled by this old-fashioned notation: volume forms are seldom exact! The volume form M is highly dependent

8.3. VOLUME FORMS

97

on the embedding of M into R N . It changes if we dilate or shrink or otherwise


deform M.
8.11. E XAMPLE . Let U be an open subset of Rn . Recall from Example 6.5
that U is a manifold covered by a single embedding, namely the identity map
: U U, x x. Then det D T D 1, so the volume form on U is simply
dt1 dt2  dtn , the ordinary volume form on Rn .
8.12. E XAMPLE . Let I be an interval in the real line and f : I R a smooth
function. Let M R2 be the graph of f . By Example 6.7 M is a 1-manifold in R 2 .
Indeed, M is the image of the embedding : I R2 given by t 5 t, f t . Let
us give M the orientation induced by the embedding , i.e. from left to right.
What is the element of arc length of M? Let us compute the pullback , a 1-form
on I. We have
1
1
1 f t 2,
D t
,
D t T D t 0 1 f t
f t
f t
so

det D t T D t dt

f t 2 dt.

The next result can be regarded as an alternative definition of M . It is perhaps


more intuitive, but it requires familiarity with Section 7.2.
8.13. P ROPOSITION . Let M be an oriented n-manifold in R N . Let x
v2 , . . . , vn Tx M. Then the volume form of M is given by

M and v1 ,

M,x v1 , v2 , . . . , vn

voln v1 , v2 , . . . , vn
voln v1 , v2 , . . . , vn
0

if v1 , v2 , . . . , vn are positively oriented,


if v1 , v2 , . . . , vn are negatively oriented,
if v1 , v2 , . . . , vn are linearly dependent,

i.e. M,x v1 , v2 , . . . , vn is the oriented volume of the n-dimensional parallelepiped in


Tx M spanned by v1 , v2 , . . . , vn .
P ROOF. For each x in M and n-tuple of tangent vectors v1 , v2 , . . . , vn at x let
x v1 , v2 , . . . , vn be the oriented volume of the block spanned by these n vectors.
This defines an n-form on M and we must show that M . Let U be an
open subset of Rn and : U R N an orientation-preserving embedding with
U M and t x for some t in U. Let us calculate the n-form on U.
We have g dt1 dt2  dtn for some function g. By Lemma 7.12 this function
is given by
g t

x e1 , e2 , . . . , en & x D t e1 , D t e2 , . . . , D t en ,

where in the second equality we used the definition of pullback. The vectors
D t e1 , D t e2 , . . . , D t en are a positively oriented basis of Tx M and, moreover, are the columns of the matrix D t , so by Theorem 8.3 they span a positive
volume of magnitude det D t T D t . This shows that g det D T D
and therefore

det D T D dt1 dt2  dtn .

Thus is equal to the local representative of M with respect to the embedding


. Since this holds for all embeddings , we have M .
QED

98

8. VOLUME FORMS

Volume form of a hypersurface. For oriented hypersurfaces M in R n there is


a more convenient expression for the volume form M . Recall the vector-valued
forms


dx1
dx1
dx ...
dx U ...
and

dxn

dxn

introduced in Section 2.5. Let n be the positive unit normal vector field on M
and let F be any vector field on M, i.e. a smooth map F : M
R n . Then the
inner product F n is a function defined on M. It measures the component of F
orthogonal to M. The product F n M is an n 1-form on M. On the other hand
we have the n 1-form F dx & F dx.
8.14. T HEOREM . On the hypersurface M we have
F dx

5 F n M .

F IRST PROOF. This proof is short but requires familiarity with the material in
Section 7.2. Let x M. Let us change the coordinates on R n in such a way that the
first n 1 standard basis vectors e1 , e2 . . . , en 1 form a positively oriented basis
of Tx M. Then, according to Example 8.7, the positive unit normal at x is given by
n x 1 n 1en and the volume form satisfies M,x e1 , . . . , en 1 1. Writing
F ni 1 Fi ei , we have F x n x &N 1 n 1 Fn x . On the other hand
F  dx

i 1

i
Fi dx1 dx



dxn ,

5 1 n 1 Fn . This proves that

F dx x e1 , . . . , en 1 &N F x n x  M e1 , . . . , en 1 ,
which implies F dx x U F x  n x M . Since this equality holds for every
x M, we find F dx 5 F n M .
QED
S ECOND PROOF. Choose an embedding : U Rn , where U is open in Rn 1 ,
such that U V M, x U . Let t U be the point satisfying t V x. As
and therefore F dx e1 , . . . , en

a preliminary step in the proof we are going to replace the embedding with a
new one enjoying a particularly nice property. Let us change the coordinates on
Rn in such a way that the first n 1 standard basis vectors e1 , e2 . . . , en 1 form
a positively oriented basis of Tx M. Then at x the positive unit normal is given by
n x N 1 n 1en . Since the columns of the Jacobi matrix D t are independent,
there exist unique vectors a1 , a2 , . . . , an 1 in Rn 1 such that D t ai ei for i 1,
2, . . . , n 1. These vectors ai are independent, because the ei are independent.
Therefore the n 1 n 1 -matrix A with i-th column vector equal to a i is
invertible. Put U A 1 U , t A 1 t and A. Then U is open in Rn 1 ,
t & x, : U Rn is an embedding with U & U , and
D t

D t DA t

D t A

by the chain rule. Therefore the i-th column vector of D t is


D t ei

D t Aei

D t a i

ei

(8.2)

8.3. VOLUME FORMS

99

for i 1, 2, . . . , n 1. (On the left ei denotes the i-th standard basis vector in
Rn 1 , on the right it denotes the i-th standard basis vector in Rn .) In other words,
the Jacobi matrix of at t is the n 1 n-matrix
I
D t & n 1 ,
0
where In 1 is the n 1  n 1 identity matrix and 0 denotes a row consisting
of n 1 zeros.
Let us now calculate F n M and F dx at the point t. Writing
F n ni 1 Fi ni and using the definition of M we get
n

F n M Fi ni a det D T D dt1 dt2  dtn 1.


i 1
From formula (8.2) we have det D t T D t 1. So evaluating this expression
at the point t and using n x &N 1 n 1en we get
F n M N 1 n 1 Fn x dt1 dt2  dtn 1 .
t

From F  dx

dx1 dx2

F dx

1 i

ni

i 1F
i
n

6dx
i 
1

i 1

From formula (8.2) we see i t t j


for 1 j n 1. Therefore

dxn we get

Fi d 1 d 2 d i  d n .
i, j for 1

i, j

n 1 and n t t j

t . 1 n 1 Fn x dt1 dt2  dtn 1 .


We conclude that 1 F n M t c F  dx t , in other words F n M x
F dx x. Since this holds for all x M we have F  dx 5 F n M .
QED
F 

dx

This theorem gives insight into the physical interpretation of n 1-forms.


Think of the vector field F as representing the flow of a fluid or gas. The direction of the vector F indicates the direction of the flow and its magnitude measures
the strength of the flow. Then Theorem 8.14 says that the n 1-form F ; dx measures, for any unit vector n in Rn , the amount of fluid per unit of time passing
through a hyperplane of unit volume perpendicular to n. We call F  dx the flux
of the vector field F.
Another application of the theorem is the following formula for the volume
form on a hypersurface. The formula provides a heuristic interpretation of the
vector-valued form dx: if n is a unit vector in Rn , then the scalar-valued n 1form n dx measures the volume of an infinitesimal n 1-dimensional parallelepiped perpendicular to n.
8.15. C OROLLARY. Let n be the unit normal vector field and M the volume form of
an oriented hypersurface M in Rn . Then

M
P ROOF. Set F

n dx.

n in Proposition 8.14. Then F n

1 because n

1. QED

100

8. VOLUME FORMS

8.16. E XAMPLE . Suppose the hypersurface M is given by an equation x s


c, where c is a regular value of a function : U R, with U open in R n . Then by
Proposition 8.8 M has a unit normal n grad grad . The volume form is
therefore n grad ; 1 grad ; dx. In particular, if M is the sphere of radius
R about the origin in Rn , then n x & x R, so M R 1 x  dx.
Exercises
8.1. Deduce from Theorem 8.3 that the area of the parallelogram spanned by a pair of
vectors a, b in Rn is given by a  b sin , where is the angle between a and b (which is
taken to lie between 0 and ). Show that a r b sin a b in R 3 .
8.2. Check that the function voln a1 , a2 , . . . , an
Definition 8.1.

  

det A T A satisfies the axioms of

8.3. Let u1 , u2 , . . . , uk and v1 , v2 , . . . , vl be vectors in R N satisfying ui v j


2, . . . , k and j 1, 2, . . . , l. (The us are perpendicular to the vs.) Prove that
volk

u1 , u2 , . . . , uk , v1 , v2 , . . . , vl

u1

0  
1 
..  ,
0. 

u2

a1

 1  a and let

0 

 a 
0 
 a 


1 . 
.. , u
.
u


c
1. 
 .a 



 1 

 a aa
 a a



2
1

a1 a2
1 a22
a3 a2
..
.
an a2

2 1

3 1

..
.
an a1

n
2
i 1 i

...,

n 1

an

be vectors in Rn 1 .
(i) Deduce from Exercise 8.3 that
(ii) Prove that

1,

a2

voln u1 , u2 , . . . , un

volk u1 , u2 , . . . , uk voll v1 , v2 , . . . , vl .

8.4. Let a1 , a2 , . . . , an be real numbers, let c

1  
0 
..  ,
0. 

0 for i

voln

a1 a3
a2 a3
1 a23
..
.
an a3

u 1 , u2 , . . . , u n , u n

...
...
...
..
.
...

a1 an
a2 an
a3 an
..
.
1 a2n





 .
1

 1  a .


n

i 1

2
i

8.5. Justify the following identities concerning orientations of a vector space V. Here
the vs form a basis of V (which in part (i) is n-dimensional and in parts (ii)(iii) twodimensional).
(i) If Sn is any permutation, then

 v   , v   , . . . , v   sign  v , v , . . . , v  .


(ii)   v , v    v , v  .
(iii)  3v , 5v   v , v  .
8.6. Let U be open in R and let f : U  R be a smooth function. Let : U  R  be
the embedding x  x, f x  and let M U  , the graph of f . Define an orientation on
M by requiring to be orientation-preserving. Deduce from Exercise 8.4 that the volume
form of M is given by   1  grad f x  dx dx   dx .
1

n 1

8.7. Let M

graph f be the oriented hypersurface of Exercise 8.6.

EXERCISES

101

*++ f - x .//
++ f - x //
+ .. / .
%
1$
n! &

"
#
1 ')( grad f x $( ,
" f - .x 0
1
&
(ii) Derive the formula 1 !
1 '2( grad f x $3( dx dx
dx from Corollary
! f " x , x , . . . , x" $ . (Caution:4 44for consistency you
8.15 by substituting x %
must replace n with n ' 1 in Corollary 8.15.)
8.8. Show that the unit normal vector field n : M 5 R defined in the proof of Proposition 8.6 is smooth. (Compute n in terms of an orientation-preserving parametrization
: U 5 M of an open subset of M.)
8.9. Let : a, b $65 R be an embedding. Let be the element of arc length on the
"
embedded
curve M ! a, b $ . Show that 1 is the 1-form on a, b $ given by ( 7 t $3( dt !
8
"
"
"
7 t $ ' 7 t $ ' 4 44 ' 7 t $ dt.
"
"
"
(i) Show that the positive unit normal vector field on M is given by
1
2

n 1

n 1

CHAPTER 9

Integration and Stokes theorem on manifolds


In this chapter we will see how to integrate an n-form over an oriented nmanifold. In particular, by integrating the volume form we find the volume of the
manifold. We will also discuss a version of Stokes theorem for manifolds. This
requires the slightly more general notion of a manifold with boundary.
9.1. Manifolds with boundary
The notion of a spherical earth developed in classical Greece around the time
of Plato and Aristotle. Older cultures (and also Western culture until the rediscovery of Greek astronomy in the late Middle Ages) visualized the earth as a flat
disc surrounded by an ocean or a void. A closed disc is not a manifold, because
no neighbourhood of a point on the edge is the image of an open subset of R 2 under an embedding. Rather, it is a manifold with boundary, a notion which can be
defined as follows. The n-dimensional halfspace is

9 : x < R = x > 0? .
;
9@: x < R = x 9 0 ? 9

Hn

9B: <

= C 0? .

Hn

The boundary of
is
int Hn
x Rn x n

Hn

Rn

and its interior is

9.1. D EFINITION . An n-dimensional manifold with boundary (or n-manifold with


boundary) in R N is a subset M of R N such that for all x M there exist

D
D
D

<

an open subset V R N containing x,


an open subset U Rn ,
and an embedding : U
R N satisfying U

9 GI

G H

Hn

I 9 VH

M.

You should compare this definition carefully with Definition 6.4 of a manifold. If
x
t with t Hn , then x is a boundary point of M. The boundary of M is the
set of all boundary points and is denoted by M. Its complement M M is the
interior of M and is denoted by int M.

<

Somewhat confusingly, the boundary of a manifold with boundary is allowed


to be empty. If nonempty, the boundary M is an n 1-dimensional manifold.
Likewise the interior int M is an n-manifold.
The most obvious example of an n-manifold with boundary is the halfspace
Hn itself, which has boundary Hn
Rn 1 and interior the open halfspace x
n
R
xn
0 . Here is a more interesting type of example, which generalizes the
graph of a function.

= C ?

: <

9 K KNM

KLF

R be a
9.2. E XAMPLE . Let U be an open subset of Rn 1 and let f : U
smooth function. Put U U
R and write elements of U as xy with x in U and
103

OP

104

9. INTEGRATION AND STOKES THEOREM ON MANIFOLDS

y in R. The region below the graph of f is the set consisting of all


y
f x .
y
M graph f

S T U

QR
x
y

in U such that

x
We assert that the region below the graph is an n-manifold whose boundary is
exactly the graph of f . We will prove this by describing it as the image of a single
Rn by
embedding. Define : U

X ut Y[Z X f T tU]t \ uY

As in Example 6.3 one verifies that is an embedding, using the fact that

X ut Y[Z X DI f^ T tU_\ 01Y ,


where 0 is the origin in R ^ . By definition therefore, the set M
Z T U ` H U is
an n-manifold in R with boundary M T U ` H U . What are M and M? A
Z
point Q R is in M if and only if it is of the form
X xyY[Z X ut Y[Z X f T tU]t \ uY
for some Q R in U ` H . Since H is given by u a 0, this is equivalent to x b U c
and y S f T x U . Thus M is exactly the region below the graph. On H we have
u 0, so M is given by the equality y
Z
Z f T xU , i.e. M is the graph.
9.3. E
. If f : U cdW R is a vector-valued map one cannot speak about
the region below the graph, but one can do the following. Again put U U cfe
Z
R. Let N n g m \ 1 and think of R as the set of vectors Q R with x in R ^ and
Z
y in R . Define : U W R by
t
X
X f T tU]\ t ue Y
u Y[Z
n 1

n 1

x
y

t
u

XAMPLE

x
y

n 1

This time we have

T U Z X DI f^ T tUh\ 0e Y
and again is an embedding. Therefore M T U ` H U is an n-manifold in R
Z
with boundary M T U ` H U . This time M is the set of points Q R of the form
Z
X yxY Z X f T tU]\ t ue Y
D t

n 1

x
y

9.1. MANIFOLDS WITH BOUNDARY

i j

l m where x is in U j and where


r o f r p xq , y s f p xq .

with t U and u 0. Hence M is the set of points


y satisfies m 1 equalities and one inequality:

x
y

o f p xq , . . . , y
Again M is given by y o f p x q , so M is the graph of f .
y1

o f p xq ,

105

y2

m 1

m 1

Here is an extension of the regular value theorem, Theorem 6.10, to manifolds


with boundary. The proof, which we will not spell out, is similar to that of Theorem 6.10.
9.4. T HEOREM (regular value theorem for manifolds with boundary). Let U be
open in R N and let : U
Rm be a smooth map. Let M be the set of x in R N satisfying

p qo

p qs c.
Suppose that c oup c , c , . . . , c q is a regular value of and that M is nonempty. Then
M is a manifold in R of codimension m n 1 and with boundary M o r p c q .
9.5. E
. Let U o R , m o 1 and p x q owv x v . The set given by the
inequality p x q s 1 is then the closed unit ball x x i R y v x v s 1 z . Since
grad p x q o 2x, any nonzero value is a regular value of . Hence the ball is an
n-manifold in R , whose boundary is r p 1 q , the unit sphere S r .
1 x

c1 ,

p qo

c2 ,

2 x

1 2
N

...,

r p xq o

cm

1,

m x

XAMPLE

n 1

If more than one inequality is involved, singularities often arise. A simple


example is the closed quadrant in R2 given by the pair of inequalities x
0 and
y 0. This is not a manifold with boundary because its edge has a sharp angle at
the origin. Similarly, a closed square is not a manifold with boundary.
However, one can show that a set given by a pair of inequalities of the form
a
f x
b, where a and b are both regular values of a function f , is a manifold
with boundary. For instance, the spherical shell

s p qs

x xi

y s v xv s

Rn R 1

R2

is an n-manifold whose boundary is a union of two concentric spheres.


Other examples of manifolds with boundary are the pair of pants, a 2-manifold
whose boundary consists of three closed curves,

and the Mbius band shown in Chapter 1. The Mbius band is a nonorientable
manifold with boundary. We will not give a proof of this fact, but you can convince

106

9. INTEGRATION AND STOKES THEOREM ON MANIFOLDS

yourself that it is true by trying to paint the two sides of a Mbius band in different
colours.
An n-manifold with boundary contained in Rn (i.e. of codimension 0) is often
called a domain. For instance, a closed ball is a domain in R n .
To define the tangent space to a manifold with boundary M at a point x choose
U and as in the definition and put

Tx M

| }~| }

D t R n .

As in the case of a manifold, this does not depend on the choice of the embedding
. Now suppose x is a boundary point of M and let v Tx M be a tangent vector.
Then v
D t u for some u Rn . We say that v points inwards if u n
0 and
outwards if un 0. If un 0, then v is tangent to the boundary. In other words,

|}

Tx M

| }~| } .

D t Rn

The above picture of the pair of pants shows some tangent vectors at boundary
points that are tangent to the boundary or outward-pointing.
Orienting the boundary. Let M be an oriented manifold with boundary. The
orientation on M induces an orientation on M by a method very similar to the
Tx M to be the unique
proof of Proposition 8.6. Namely, for x M define n x
outward-pointing tangent vector of length 1 which is orthogonal to Tx M. This defines the unit outward-pointing normal vector field on M. A basis v1 , v2 , . . . , vn 1
of Tx M is called positively oriented if n x , v1 , v2 , . . . , vn 1 is a positively oriented basis of Tx M. This defines an orientation of M, called the induced orientation.
For instance, let M Hn with the standard orientation e1 , . . . , en . At each point
of M
Rn 1 the outward pointing normal is en . This implies that induced
orientation on M is
1 n e1 , e2 , . . . , en 1 , because

| }

| | }

| }


e , e , e , . . . , e {| 1 } e , e , . . . , e
n

n 1

n 1 , en

9.2. Integration over orientable manifolds


As we saw in Chapter 5, a form of degree n can be integrated over a chain
of dimension n. The integral does not change if we reparametrize the chain in
an orientation-preserving manner. This opens up the possibility of integrating an
n-form over an oriented n-manifold.
Let M be an n-dimensional oriented manifold (possibly with boundary) in R N
and let be an n-form on M. To define the integral of over M let us assume
that M is compact. (A subset of R N is called compact if it is closed and bounded;
see Appendix A.2. This assumption is made to ensure that the integral is a proper
integral and therefore converges.) For a start, let us also make the assumption that
there exists a smooth map c : 0, 1 n
R N such that


c 0, 1 { M and
the restriction of c to | 0, 1 } is an orientation-preserving embedding.
For instance, this assumption is satisfied for the n-sphere S (see Exercise 5.5) and
the torus S S (see Exercise 9.4). The pullback c is then an n-form on the cube
0, 1 . We define

{
c .
n

0,1

9.2. INTEGRATION OVER ORIENTABLE MANIFOLDS

107

Suppose c : 0, 1 n
R N is a smooth map with the same properties as c. To ensure
that M is well-defined we need to check the following equality.

c c .
S
. Let us denote the closed cube 0, 1 by R and let U be
the open cube 0, 1 . Let V U c c U and V U c c U . The

9.6. L EMMA .

0,1

0,1

KETCH OF PROOF
n

complement of V and of V in R are negligeable in the sense that

c c
R

c c .

and

(9.1)

By assumption the restriction of c to U is an embedding. This implies that c : V


M is a bijection onto its image, and so we see that c 1 c is a bijection from V
onto V. It is orientation-preserving, because c and c are orientation-preserving.
Therefore, by Theorem 5.1,

c c c c c c c c .
1

Combining this with the equalities (9.1) we get the result.

QED

Not every manifold can be covered with one single n-cube. However, it can
be shown that there always exists a finite collection of n-cubes c i : 0, 1 n
M for
i 1, 2, . . . , k, such that

M,
(i) ki 1 ci 0, 1 n
n
(ii) ci 0, 1
c j 0, 1 n is empty for i
j,
(iii) for each i the restriction of c i to 0, 1 n is an orientation-preserving embedding.

We can then define

i 1

c ,
0,1

and check as in Lemma 9.6 that the result does not depend on the maps c i . (The
condition (ii) on the maps is imposed to avoid double counting in the integral.)
9.7. D EFINITION . Let M a compact oriented manifold in R N . The volume of M
is vol M
1, resp. 2, we
M , where is the volume form on M. (If dim M
speak of the arc length, resp. surface area of M.) The integral of a function f on M
is defined as M f . The mean or average of f is the number f
vol M 1 M f .
n
The centroid or barycentre of M is the point x in R whose i-th coordinate is the
mean value of xi over M, i.e.

x i

1
vol M

xi .

The volume form depends on the embedding of M into R N , so the notions


defined above depend on the embedding as well.
The most important property of the integral is the following version of Stokes
theorem, which can be viewed as a parametrization-independent version of Theorem 5.10 and is proved in a similar way.

108

9. INTEGRATION AND STOKES THEOREM ON MANIFOLDS

9.8. T HEOREM (Stokes theorem for manifolds). Let be an n 1-form on a


compact oriented n-manifold with boundary M. Give the boundary M the induced orientation. Then

9.3. Gau and Stokes


Stokes theorem, Theorem 9.8, contains as special cases the integral theorems
of vector calculus. These classical results involve a vector field F
ni 1 Fi ei den
fined on an open subset U of R . As discussed in Section 2.5, to this vector field
corresponds a 1-form F dx ni 1 Fi dxi , which we can think of as the work
done by the force F along an infinitesimal line segment dx. We will now derive
the classical integral theorems by applying Theorem 9.8 to one-dimensional, resp.
n-dimensional, resp. two-dimensional manifolds M contained in U.

Fundamental theorem of calculus. If F is conservative, F grad g for a function g, then


grad g dx
dg. If M is a compact oriented 1-manifold with
boundary in Rn , then M dg
M g by Theorem 9.8. The boundary consists of
two points a and b (if M is connected). If the orientation of M is from a to b,
then a acquires a minus and b a plus. Stokes theorem therefore gives the fundamental theorem of calculus in Rn ,

g b g a .

F dx

If we interpret F as a force acting on a particle travelling along M, then g stands


for the potential energy of the particle in the force field. Thus the potential energy
of the particle decreases by the amount of work done.
Gau divergence theorem. We have

F dx

and

d
n

div F dx 1 dx2

dx .

If N is a oriented hypersurface in R with positive unit normal n, then


F
n N on N by Theorem 8.14. In this situation it is best to think of F as the flow
vector field of a fluid, where the direction of F x gives the direction of the flow at
a point x and the magnitude F x gives the mass of the amount of fluid passing
per unit time through a hypersurface of unit area placed at x perpendicular to the
vector F x . Then describes the amount of fluid passing per unit time and per
unit area through the hypersurface N. For this reason the n 1-form is also
called the flux of F, and its integral over N the total flux through N.
Applying Stokes theorem to a compact domain M in Rn we get M d
M . Written in terms of the vector field F this is Gau divergence theorem,

div F dx1 dx2

dx
n

F n

M .

Thus the total flux out of the hypersurface M is the integral of div F over M. If
the fluid is incompressible (e.g. most liquids) then this formula leads to the interpretation of the divergence of F (or equivalently d ) as a measure of the sources
or sinks of the flow. Thus div F
0 for an incompressible fluid without sources

EXERCISES

109

or sinks. If the fluid is a gas and if there are no sources or sinks then div F x
(resp. 0) indicates that the gas is expanding (resp. being compressed) at x.

Classical version of Stokes theorem. Now let M be a compact two-dimensional oriented surface with boundary and let us rewrite Stokes theorem M d
M in terms of the vector field F. The right-hand side represents the work of
F done around the boundary curve(s) of M, which is not necessarily 0 if F is not
conservative. The left-hand side has a nice interpretation if n
3. Then d
curl F dx, so d curl F dx. Hence if n is the positive unit normal of the surface
M in R3 , then d
curl F n M on M. In this way we get the classical formula
of Stokes,

curl F n

F dx.

In other words, the total flux of curl F through the surface M is equal to the work
done by F around the boundary curves of M. This formula shows that curl F, or
equivalently d , can be regarded as a measure of the vorticity of the vector field.

Exercises




(i) Draw a picture of M if U is the open unit disc given by x y
1 x y and g x, y 2 x y .

R be two smooth functions


9.1. Let U be an open subset of Rn and let f , g : U
g x for all x in U. Let M be the set of all pairs x, y such that x in U and
satisfying f x
f x
y g x .
2

1 and f x, y

(ii) Show directly from the definition that M is a manifold with boundary. (Use
two embeddings to cover M.) What is the dimension of M and what are the
boundary and the interior?
(iii) Give an example showing that M is not necessarily a manifold with boundary if
the condition f x
y g x fails.


9.2.
(i) Let x dy y dx and let M be a compact domain in the plane R .
Show that
is twice the surface area of M.
(ii) Apply the observation of part (i) to find the area enclosed by the astroid x
cos t, y sin t.
(iii) Let x dx and let M be a compact domain in R . Show that
is a
constant times the volume of M. What is the value of the constant?
9.3. Write the divergence theorem for the vector field F cx e on R , where c is
2

n n

a positive constant. Deduce Archimedes Law: the buoyant force exerted on a submerged
body is equal to the weight of the displaced fluid. E

!
by

9.4. Let R1

R2

0 be constants. Define a 3-cube c : 0, R 2

r
1
2

R
R
1

r cos 2 cos 1
r cos 2 sin 1
r sin 2

N 0, 2 0, 2

R3

(i) Sketch the image of c.


(ii) Let x 1 , x2 , x3 be the standard coordinates on R3 . Compute c dx1 , c dx2 , c dx3
and c dx1 dx2 dx3 .
(iii) Find the volume of the solid parametrized by c.
(iv) Find the surface area of the boundary of this solid.

110

9. INTEGRATION AND STOKES THEOREM ON MANIFOLDS

9.5. Let M be a compact domain in Rn . Let f and g be smooth functions on M. The


Dirichlet integral of f and g is D f , g
dx 1 dx2
dxn is
M grad f grad g , where
the volume form on M.
(i) Show that d f dg
grad f grad g .
(ii) Show that d dg
g , where g ni 1 2 g x2i .
(iii) Deduce from parts (i)(ii) that d f dg
grad f grad g f g .
(iv) Let n be the outward-pointing unit normal vector field on M. Write g n for
the directional derivative Dg n grad g n. Show that

] 




g
f dg
f
.
n
(v) Deduce from parts (iii) and (iv) Greens formula,
g

D f , g
f

f g .
n
M

g
n

(vi) Deduce Greens symmetric formula,


f

n M

f g g f .

9.6. In this problem we will calculate the volume of a ball and a sphere in Euclidean
space. Let B R be the closed ball of radius R about the origin in R n . Then its boundary
S R
B R is the sphere of radius R. Put Vn R
voln B R and An R
voln 1 S R .
Also put Vn Vn 1 and An
An 1 .
(i) Deduce from Corollary 8.15 that the volume form on S R is the restriction of
to S R , where is as in Exercise 2.16. Conclude that A n R
S R .
Rn Vn and An R
Rn 1 An . (Substitute y
Rx in the
(ii) Show that Vn R
volume forms of B R and S R .)
(iii) Let f : 0,
R be a continuous function. Define g : R n
R by g x
f x . Use Exercise 2.16(ii) to prove that

g dx dx  dx
f r A r dr A
f r r dr.

(iv) Show that



e
A
dr
dr.
e r

in part (iii) and let R .)


(Take f r e
R

B R

r2

n 1

r2 n 1

r2

(v) Using Exercises B.10 and B.11 conclude that

2
2 
.
and A 

m
1
!
1
3
5   2m 1

1 in part (iii) show that A nV and A R V R R.


(vi) By taking f r

An

2 2



n
2

whence



2
n
2

1

whence

2m 1

(vii) Deduce that


Vn

m 1

A 2m

V2m

m
m!

V2m 

and

1 3 25   2m 1 .
m 1 m

(viii) Complete the following table. (Conventions: a space of negative dimension is


empty; the volume of a zero-dimensional manifold is its number of points.)
n

Vn R
An R

R2

4
3
3 R

2 R

EXERCISES

(ix) Find limn 

111

An , limn  Vn and limn 


An
lim



x

xx

1
1
2

ex



2 .

An . Use Stirlings formula,

CHAPTER 10

Applications to topology
10.1. Brouwers fixed point theorem
Let M be a manifold, possibly with boundary. A retraction of M onto a subset
A is a smooth map : M  A such that  x  x for all x in A. For instance, let
M be the punctured unit ball in n-space,
M  x  Rn  0  x 

1! .

Then the normalization map  x " x #$ x  is a retraction of M onto its boundary
A  M, the unit sphere. The following theorem says that a retraction onto the
boundary is impossible if M is compact and orientable.
10.1. T HEOREM . Let M be a compact orientable manifold with nonempty boundary.
Then there does not exist a retraction from M onto M.
P ROOF. Suppose : M  M was a retraction. Let us choose an orientation
of M and equip M with the induced orientation. Let  M be the volume form
on the boundary (relative to some embedding of M into R N ). Let  % be its
pullback to M. Let n denote the dimension of M. Note that is an n & 1-form
on the n & 1-manifold M, so d  0. Therefore d  d %  % d  0 and
hence by Stokes theorem 0 (' M d )' M . But is a retraction onto M, so the
restriction of to M is the identity map and therefore  on M. Thus
0 )*

)*

vol M  +

which is a contradiction. Therefore does not exist.

0,
QED

This brings us to one of the oldest results in topology. Suppose f is a map from
a set X into itself. An element x of X is a fixed point of f if f  x , x.
10.2. T HEOREM (Brouwers fixed point theorem). Every smooth map from the
closed unit ball into itself has at least one fixed point.
P ROOF. Let M - x  Rn   x . 1 ! be the closed unit ball. Suppose
f : M  M was a smooth map without fixed points. Then f  x / + x for all x. For
each x in the ball consider the halfline starting at f  x  and pointing in the direction
of x. This halfline intersects the unit sphere M in a unique point that we shall call
113

114

10. APPLICATIONS TO TOPOLOGY

0 x 1 , as in the following picture.


f 2 x3

2 y3

x
y
f 2 y3

2 x3

This defines a smooth map : M 4 M. If x is in the unit sphere, then 0 x 1"5 x,


so is a retraction of the ball onto its boundary, which contradicts Theorem 10.1.
Therefore f must have a fixed point.
QED
This theorem can be stated imprecisely as saying that after you stir a cup of
coffee, at least one molecule must return to its original position. Brouwer originally stated his result for arbitrary continuous maps. This more general statement
can be derived from Theorem 10.2 by an argument from analysis which shows
that every continuous map is homotopic to a smooth map. (See Section 10.2 for
the definition of homotopy.) The theorem also remains valid if the closed ball is
replaced by a closed cube or a similar shape.
10.2. Homotopy
Definition and first examples. Suppose that 0 and 1 are two maps from a
manifold M to a manifold N and that is a form on N. What is the relationship
between the pullbacks 06 and 16 ? There is a reasonable answer to this question
if 0 can be smoothly deformed into 1 . More formally, we say that 0 and 1
are homotopic if there exists a smooth map : M 78 0, 1 9:4 N such that 0 x, 0 1"5
0 0 x 1 and 0 x, 1 1;5 1 0 x 1 for all x in M. The map is called a homotopy. Instead of
0 x, t 1 we often write t 0 x 1 . Then each t is a map from M to N and we can think
of t as a family of maps parametrized by t in the unit interval that interpolates
between 0 and 1 , or as a one-second movie that at time 0 starts at 0 and at
time 1 ends up at 1 .
10.3. E XAMPLE . Let M 5 N 5 Rn and 0 0 x 1<5 x (identity map) and 1 0 x 1,5 0
(constant map). Then 0 and 1 are homotopic. A homotopy is given by 0 x, t 1,5
0 1 = t 1 x. This homotopy collapses Euclidean space onto the origin by moving
each point radially inward. There are other ways to accomplish this. For instance
0 1 = t 1 2 x and 0 1 = t2 1 x are two other homotopies between the same maps. We can
also interchange 0 and 1 : if 0 0 x 1<5 0 and 1 0 x 1,5 x, then we find a homotopy
by reversing time (playing the movie backwards), 0 x, t 1,5 tx.
10.4. E XAMPLE . Let M 5 N be the punctured Euclidean space R n =?> 0 @ and
let 0 0 x 1A5 x (identity map) and 1 0 x 1B5 x C$D x D (normalization map). Then 0
and 1 are homotopic. A homotopy is given for instance by 0 x, t 1E5 x C$D x D t or
by 0 x, t 1/5F0 1 = t 1 x G tx C$D x D . Either of these homotopies collapses punctured
Euclidean space onto the unit sphere about the origin by smoothly stretching or
shrinking each vector until it has length 1.

10.2. HOMOTOPY

115

10.5. E XAMPLE . A manifold M is said to be contractible if there exists a point


x0 in M such that the constant map 0 H x IKJ x0 is homotopic to the identity map
H x I<J x. A specific homotopy : M LNM 0, 1 O$P M from 0 to 1 is a contraction of
M onto x0 . (Perhaps expansion would be a more accurate term, a contraction
being the result of replacing t with 1 Q t.) Example 10.3 shows that R n is contractible onto the origin. (In fact it is contractible onto any point x 0 . Can you write
a contraction of Rn onto x0 ?) The same formula shows that an open or closed ball
around the origin is contractible. We shall see in Theorem 10.19 that punctured
n-space Rn Q?R 0 S is not contractible.
Homotopy of curves. If M is an interval M a, b O and N any manifold, then maps
from M to N are nothing but parametrized curves in N. A homotopy of curves can
be visualized as a piece of string moving through the manifold N.

Homotopy of loops. A loop in a manifold N is a smooth map from the unit


circle S1 into N. This can be visualized as a thin rubber band sitting in N. A
homotopy of loops : S1 LM 0, 1 O;P N can be pictured as a rubber band floating
through N from time 0 until time 1.

N
S

10.6. E XAMPLE . Consider the two loops 0 , 1 : S1 P R2 in the plane given


by 0 H x ITJ x and 1 H x IUJ x VXW 20 Y . A homotopy of loops is given by shifting
0 to the right, t H x IZJ x V[W 2t
0 Y . What if we regard 0 and 1 as loops in the
2
punctured plane R Q.R 0 S ? Clearly the homotopy does not work, because it
moves the loop through the forbidden point 0. (E.g. t H x IEJ 0 for x J W]\ 10 Y and
t J 1 ^ 2.) In fact, however you try to move 0 to 1 you get stuck at the origin, so

116

10. APPLICATIONS TO TOPOLOGY

it seems intuitively clear that there exists no homotopy of loops from 0 to 1 in


the punctured plane. This is indeed the case, as we shall see in Example 10.13.
The homotopy formula. The product M _a` 0, 1b is often called the cylinder
with base M. The two maps defined by 0 c x d,e c x, 0 d and 1 c x d<e c x, 1 d send M to
the bottom, resp. the top of the cylinder. A homotopy : M _` 0, 1b,f M _` 0, 1 b
between these maps is given by the identity map c x, t d,e c x, t d . (Slide the bottom
to the top at speed 1.)

1
4 g

10

0
base

cylinder

If M is an open subset of Rn , a k h 1-form on the cylinder can be written as

f I c x, t d
I

dx I h

g J c x, t d
J

dt dx J ,

with I running over multi-indices of degree k h 1 and J over multi-indices of degree k. (Here we write the dt in front of the dxs because that is more convenient in
what follows.) The cylinder operator turns forms on the cylinder into forms on the
base lowering the degree by 1,

: i

kj 1

c M _k` 0, 1 bldmfni

c Md ,

by taking the piece of involving dt and integrating it over the unit interval,

op

g J c x, t d dt q dx J .

(In particular e 0 for any that does not involve dt.) For a general manifold
M we can write a k h 1-form on the cylinder as e h dt , where and are
forms on M _` 0, 1 b (of degree k h 1 and k respectively) that do not involve dt. We
then define e(r 01 dt.
The following result will enable us to compare pullbacks of forms under homotopic maps. It can be regarded as an application of Stokes theorem, but we
shall give a direct proof.
10.7. L EMMA (cylinder formula). Let M be a manifold. Then 1s t 0s e d h
d for all k h 1-forms on M _k` 0, 1b . In short,

1s t 0s e d h d .
P ROOF. We write out the proof for an open subset of Rn . The proof for arbitrary manifolds is similar. It suffices to consider two cases: e
f dx I and
e g dt dx J .

10.2. HOMOTOPY

Case 1. If u

f dx I , then u

f
dt dx I v
t

d u

0 and d u

xi dxi dx I u
i

117

0. Also

f
dt dx I v terms not involving dt,
t

so
1

d v d u d uxwy
Case 2. If u

f
x, t { dt | dx I
t z

g dt dx J , then 0 u 1 u
g

so

d u
Also u

}y

d u

0 and

xi dxi dt dx J u xi dt dxi dx J ,

d u

u~} f x, 1 { f x, 0 { dx I u 1  0 .
z
z

w y

i 1

1
0

g
x, t { dt | dxi dx J .
xi z

g x, t { dt dx J , so

xi w y

i 1

Hence d v d u

1
0

g x, t { dt | dxi dx J u

w y

i 1

1
0

g
x, t { dt | dxi dx J .
xi z

0 u 1  0 .

QED

Now suppose we have a pair of maps 0 and 1 going from a manifold M


to a manifold N and that : M 0, 1 < N is a homotopy between 0 and 1 .
For x in M we have 0 x {Tu x, 0 {Tu 0 x { , in other words 0 u 0 .
z
z
z
Similarly 1 u 1 . Hence for any k v 1-form on N we have 0 u 0 and
1 u 1 . Applying the cylinder formula to u we see that the pullbacks
0 and 1 are related in the following manner.
10.8. T HEOREM (homotopy formula). Let 0 , 1 : M N be smooth maps from
a manifold M to a manifold N and let : M  0, 1 N be a homotopy from 0 to 1 .
Then 1  0 u d v d for all k v 1-forms on N. In short,

1  0 u d v d .
In particular, if d u

0 we get 1 u

0 v d .

10.9. C OROLLARY. If 0 , 1 : M N are homotopic maps between manifolds and


is a closed form on N, then 0 and 1 differ by an exact form.
This implies that if the degree of is equal to the dimension of M, 0 and
1 have the same integral.
10.10. T HEOREM . Let M and N be manifolds and let be a closed n-form on N,
where n u dim M. Suppose M is compact and oriented and has no boundary. Let 0 and
1 be homotopic maps from M to N. Then

0 u)y

1 .

118

10. APPLICATIONS TO TOPOLOGY

P ROOF. By Corollary 10.9, 1 0


by Stokes theorem

because M is empty.

1 0 ;

d for an n 1-form on M. Hence

0,
QED

A LTERNATIVE PROOF. Here is a proof based on Stokes theorem for the manifold with boundary M  0, 1 . The boundary of M  0, 1 consists of two copies
of M, namely M  1 and M  0 , the first of which is counted with a plus
sign and the second with a minus. Therefore, if : M  0, 1 N is a homotopy

between 0 and 1 ,
0

M 0,1

M 0,1

M 0,1

0 .
QED

If M is the circle S1 , N the punctured plane R2 . 0 and the angle form


y dx x dy x2 y2 of Example 3.8, then a map from M to N is a loop in N
and the integral of is 2 times the winding number of the loop. Thus Theorem
10.10 gives the following result.
10.11. C OROLLARY. Homotopic loops in R 2 a 0 have the same winding number
about the origin.
10.12. E XAMPLE . Unfolding the three self-intersections in the curve pictured
below does not affect its winding number.

10.13. E XAMPLE . The two circles 0 and 1 of Example 10.6 have winding
number 1, resp. 0 and therefore are not homotopic (as loops in the punctured
plane).
10.3. Closed and exact forms re-examined
The homotopy formula throws light on our old problem of when a closed form
is exact, which we looked into in Section 2.3. The answer turns out to depend on
the shape of the manifold on which the forms are defined. On some manifolds
all closed forms (of positive degree) are exact, on others this is true only in certain
degrees. Failure of exactness is typically detected by integrating over a submanifold of the correct dimension and finding a nonzero answer. In a certain sense all
obstructions to exactness are of this nature. We shall not attempt to say the last

10.3. CLOSED AND EXACT FORMS RE-EXAMINED

119

word on this problem, but study a few representative special cases. The matter is
explored in [Fla89] and at a more advanced level in [BT82].
0-forms. A closed 0-form on a manifold is a smooth function f satisfying
d f 0. This means that f is constant (on each connected component of M). If
this constant is nonzero, then f is not exact (because forms of degree 1 are by
definition 0). So a closed 0-form is never exact (unless it is 0) for a rather uninteresting reason.
1-forms and simple connectivity. Let us now consider 1-forms on a manifold
M. Theorem 4.5 says that the integral of an exact 1-form along a loop is 0. With
a stronger assumption on the loop the same is true for arbitrary closed 1-forms. A
loop c : S1 M is null-homotopic if it is homotopic to a constant loop. The integral
of a 1-form along a constant loop is 0, so from Theorem 10.10 (where we set the M
of the theorem equal to S 1 ) we get the following.
10.14. P ROPOSITION . Let c be a null-homotopic loop in M. Then c
closed forms on M.

0 for all

A manifold is simply connected if every loop in it is null-homotopic.


10.15. T HEOREM . All closed 1-forms on a simply connected manifold are exact.
P ROOF. Let be a closed 1-form and c a loop in M. Then c is null-homotopic,
so c 0 by Proposition 10.14. The result now follows from Theorem 4.5. QED
10.16. E XAMPLE . The punctured plane R2 ? 0 is not simply connected, because it possesses a nonexact closed 1-form. (See Example 4.6.) In contrast it can be
proved that for n 3 the sphere S n 1 and punctured n-space Rn  0 are simply
connected. Intuitively, the reason is that in two dimensions a loop that encloses
the puncture at the origin cannot be crumpled up to a point without getting stuck
at the puncture, whereas in higher dimensions there is enough room to slide any
loop away from the puncture and then squeeze it to a point.
The Poincar lemma. On a contractible manifold all closed forms of positive
degree are exact.
10.17. T HEOREM (Poincar lemma). All closed k-forms on a contractible manifold
are exact for k 1.
P ROOF. Let M be a manifold and let : M k 0, 1 M be a contraction onto
a point x0 in M, i.e. a smooth map satisfying x, 0 m x0 and x, 1 m x for all x.
Let be a closed k-form on M with k 1. Then 1 and 0 0, so putting
we get
d

1 0 d .

Here we used the homotopy formula, Theorem 10.8, and the assumption that d
0. Hence d .
QED
The proof provides us with a formula for the antiderivative, namely
, which can be made quite explicit in certain cases.

120

10. APPLICATIONS TO TOPOLOGY

10.18. E XAMPLE . Let M be Rn and let x, t


Let i gi dxi be a 1-form. Then


so

tx be the radial contraction.

gi tx d txi ; gi tx xi dt
i

xi

t dxi ,

gi tx dt.

According to the proof of the Poincar lemma, the function satisfies d provided that d 0. It is instructive to compare with the function f constructed
in the proof of Theorem 4.5. (See Exercise 10.5.)
Another typical application of the Poincar lemma is showing that a manifold
is not contractible by exhibiting a closed form that is not exact. For example, the
punctured plane R2 0 is not contractible because it possesses a nonexact closed
1-form, namely the angle form. (See Example 4.6.) The angle form generalizes to
an n 1-form on punctured n-space Rn 0 ,

x dx
n .
x

10.19. T HEOREM . is a closed but non-exact n 1-form on punctured n-space.


Hence punctured n-space is not contractible.
P ROOF. d 0 follows from Exercise 2.1(ii). The n 1-sphere M S n 1 has
unit normal vector field x, so by Corollary 8.15 on M we have , the volume
form. Hence M vol M 0. On the other hand, suppose was exact, d
for an n 1-form . Then

by Stokes theorem, Theorem 9.8. This is a contradiction, so is not exact. It now


follows from the Poincar lemma, Theorem 10.17, that R n N 0 is not contractible.
QED
Using the same form , but restricting it to the unit sphere S n 1 , we see that
is not contractible. But how about forms of degree not equal to n 1? Without proof we state the following fact.
Sn 1

10.20. T HEOREM . On Rn a 0 and on Sn

n 1 is exact.

every closed form of degree k

1,

For a compact oriented hypersurface without boundary M contained in in


Rn ? 0 the integral
1
x dx
n

1
n
x
vol n 1 S
M
is the winding number of M about the origin. It generalizes the winding number
of a closed curve in R2 0 around the origin. It can be shown that the winding
number in any dimension is always an integer. It provides a measure of how
many times the hypersurface wraps around the origin. For instance, the proof
of Theorem 10.19 shows that the winding number of the n 1-sphere about the
origin is 1.

10.3. CLOSED AND EXACT FORMS RE-EXAMINED

121

Contractibility versus simple connectivity. Theorems 10.15 and 10.17 suggest that the notions of contractibility and simple connectivity are not independent.
10.21. P ROPOSITION . A contractible manifold is simply connected.
P ROOF. Use a contraction to collapse any loop onto a point.

x0

x0

c1
M

c1
M

Formally, let c1 : S1 M be a loop, : M 0, 1 M a contraction of M onto x0 .


Put c s, t K c1 s , t . Then c is a homotopy between c 1 and the constant loop
c0 t ; c1 s , 0 m x0 positioned at x0 .
QED
As mentioned in Example 10.16, the sphere S n 1 and punctured n-space Rn
0 are simply connected for n 3, although it follows from Theorem 10.19 that
they are not contractible. Thus simple connectivity is weaker than contractibility.

The Poincar conjecture. Not long after inventing the fundamental group
Poincar posed the following question. Let M be a compact three-dimensional
manifold without boundary. Suppose M is simply connected. Is M homeomorphic to the three-dimensional sphere? (This means: does there exist a bijective
map M S3 which is continuous and has a continuous inverse?) This question
became (inaccurately) known as the Poincar conjecture. It is famously difficult and
was the force that drove many of the developments in twentieth-century topology.
It has an n-dimensional analogue, called the generalized Poincar conjecture, which
asks whether every compact n-dimensional manifold without boundary which is
homotopy equivalent to Sn is homeomorphic to Sn . We cannot here go into this
fascinating problem in any serious way, other than to report that it has now been
completely solved. Strangely, the case n 5 of the generalized Poincar conjecture conjecture was the easiest and was confirmed by S. Smale in 1960. The case
n 4 was done by M. Freedman in 1982. The case n 3, the original version of
the conjecture, turned out to be the hardest, but was finally confirmed by G. Perelman in 2002-03. For a discussion and references, see the paper Towards the Poincar
conjecture and the classification of 3-manifolds by J. Milnor, which appeared in the
November 2003 issue of the Notices of the American Mathematical Society and
can be read online at <<]: .

122

10. APPLICATIONS TO TOPOLOGY

Exercises
10.1. Write a formula for the map figuring in the proof of Brouwers fixed point
theorem and prove that it is smooth.
10.2. Let x0 be any point in Rn . By analogy with the radial contraction onto the origin,
write a formula for radial contraction onto the point x 0 . Deduce that any open or closed
ball centred at x0 is contractible.
10.3. A subset M of Rn is star-shaped relative to a point x 0 M if for all x M the
straight line segment joining x 0 to x is entirely contained in M. Show that if M is starshaped relative to x 0 , then it is contractible onto x 0 . Give an example of a contractible set
that is not star-shaped.
10.4. A subset M of Rn is convex if for all x and y in M the straight line segment joining
x to y is entirely contained in M. Prove the following assertions.
(i) M is convex if and only if it is star-shaped relative to each of its points. Give an
example of a star-shaped set that is not convex.
(ii) The closed ball B , x of radius centred at x is convex.
(iii) Same for the open ball B , x .
10.5. Let x Rn and let cx be the straight line connecting the origin to x. Let be a
1-form on Rn and let be the function defined in Example 10.18. Show that x ? cx .

10.6. Let be the k-form f dx I f dxi1 dxi2


be the radial contraction x, t tx. Verify that

k
1 m 1

m 1

and check directly that d

f tx tk

dxik

on Rn and let : Rn

dt xim dxi1 dxi2

for k 

0, 1 

Rn

dxim dxik ,

1.

10.7. Let
f dx dy g dz dx h dy dz be a 2-form on R3 and let x, y, z, t
t x, y, z be the radial contraction of R3 onto the origin. Verify that

1
0

f tx, ty, tz t dt " x dy

y dx

1
0

g tx, ty, tz t dt

1
0

z dx

x dz

h tx, ty, tz t dt " y dz

z dy .

10.8. Let I f I dx I be a closed k-form whose coefficients f I are smooth functions


defined on Rn  0  that are all homogeneous of the same degree p
 k. Let

Show that d . (Use d


also Exercise 2.7.)

] 1 l
k

I l 1

xil f I dxi1 dxi2

dxil dxik .

0 and apply the identity proved in Exercise B.5 to each f I ; see

10.9. Let M and N be manifolds and 0 , 1 : M


N homotopic maps. Show that


for
all
closed
k-chains
c
in
M
and
all
closed
k-forms on N.

c 0
c 1

10.10. Prove that any two maps 0 and 1 from M to N are homotopic if M or N is
contractible. (First show that every map M N is homotopic to a constant map x
y0 .)
10.11. Let x0 2, 0 and let M be the twice-punctured plane R 2  0, x0  . Let c1 , c2 ,

M be the loops defined by c 1 t cos t, sin t , c 2 t ) 2 cos t, sin t and
2 cos t, 2 sin t . Show that c 1 , c2 and c3 are not homotopic. (Construct a 1-form
on M such that the integrals c1 , c2 and c3 are distinct.)
c3 : 0, 2 

c3 t 1

EXERCISES

123

10.12. A function g : R
R is 2 -periodic if g x 2  g x for all x.
(i) Let g : R
R be a smooth 2 -periodic function and let  g dt, where t is
the coordinate on R. Prove that there is a unique number k such that  k dt 
dh for some smooth 2 -periodic function h. (To find k, integrate the equation
 k dt  dh over  0, 2  . Then check that this value of k works.)
(ii) Let be any 1-form on the unit circle S 1 and let be the element of arc length
of S1 . (You can think of as the restriction to S 1 of the angle form.) Prove that
there is a unique number k such that  k is exact. (Use the parametrization
c t  cos t, sin t and apply the result of part (i).)

APPENDIX A

Sets and functions


A.1. Glossary
We start with a list of set-theoretical notations that are frequently used in the
text. Let X and Y be sets.
x  X: x is an element of X.
a, b, c  : the set containing the elements a, b and c.
X  Y: X is a subset of Y, i.e. every element of X is an element of Y.
X  Y: the intersection of X and Y. This is defined as the set of all x such
that x  X and x  Y.
X  Y: the union of X and Y. This is defined as the set of all x such that
x  X or x  Y.
X  Y: the complement of Y in X. This is defined as the set of x in X such
that x is not in Y.
X  Y: the Cartesian product of X and Y. This is by definition the set of
all ordered pairs  x, y  with x  X and y  Y. Examples: R  R is
the Euclidean plane, usually written R2 ; S1   0, 1 ! is a cylinder wall of
height 1; S1  S1 is a torus.

R2

S1 "$# 0, 1 %

S1 " S1

x  X & P  x  : the set of all x  X which have the property P  x  . Examples:




x R & 1'

x(

3  is the interval  1, 3  ,

x & x  X and x  Y  is the intersection X  Y,



x & x  X or x  Y  is the union X  Y,

x  X & x  ) Y  is the complement X  Y.

f : X * Y: f is a function (also called a map) from X to Y. This means that


f assigns to each x  X a unique element f  x + Y. The set X is called
the domain or source of f , and Y is called the codomain or target of f .
125

126

A. SETS AND FUNCTIONS

f , A - : the image of a A under the map f . If A is a subset of X, then its image


under f is by definition the set
f , A -/.10 y 2 Y 3 y .
f5

f , x - for some x 2 A 4 .

1,

B - : the preimage of B under the map f . If B is a subset of Y, this is by


definition the set
f5

, B -6.70 x 2 X 3 f , x -82 B 4 .

(This is a somewhat confusing notation. It is not meant to imply that f is


required to have an inverse.)
f 5 1 , c - : an abbreviation for f 5 1 ,0 c 4- , i.e. the set 0 x 2 X 3 f , x -9. c 4 . This
is often called the fibre or level set of f at c.
f 3 A: the restriction of f to A. If A is a subset of X, f 3 A is the function defined
by
<
f , xif x 2 A,
, f 3 A -:, x -;.
not defined if x 2 = A.
In other words, f 3 A is equal to f on A, but forgets the values of f at
points outside A.
g > f : the composition of f and g. If f : X ? Y and g : Y ? Z are functions,
then g > f : X ? Z is defined by , g > f , x -@. g , f , x -A- . We often say that
the function g > f is obtained by substituting y . f , x - into g , y - .
A function f : X ? Y is injective or one-to-one if x 1 . = x2 implies f , x1 -B. = f , x2 - .
(Equivalently, f is injective if f , x 1 -8. f , x2 - implies x1 . x2 .) It is called surjective
or onto if f , X - . Y, i.e. if y 2 Y then y . f , x - for some x 2 X. It is called
bijective if it is both injective and surjective. The function f is bijective if and only
if it has a two-sided inverse f 5 1 : Y ? X satisfying f 5 1 , f , x -A-B. x for all x 2 X
and f , f 5 1 , y -A-6. y for all y 2 Y.
If X is a finite set and f : X ? R a real-valued function, the sum of all the
numbers f , x - , where x ranges through X, is denoted by x C X f , x - . The set X is
called the index set for the sum. This notation is often abbreviated or abused in
various ways. For instance, if X is the collection 0 1, 2, . . . , n 4 , one uses the familiar
notation niD 1 f , i - . In these notes we will often deal with indices which are pairs
or k-tuples of integers, also known as multi-indices. As a simple example, let n be
a fixed nonnegative integer, let X be the set of all pairs of integers , i, j - satisfying
0 E i E j E n, and let f , i, j -6. i F j. For n . 3 we can display X and f in a tableau
as follows.
j
3

EXERCISES

The sum x G

127

f H x I of all these numbers is written as

0J iJ jJ n

HiK

jI .

You will be asked to evaluate it explicitly in Exercise A.2.


A.2. General topology of Euclidean space
Let x be a point in Euclidean space Rn . The open ball of radius about a point
x is the collection of all points y whose distance to x is less than ,
B LMH , x I6NPO y Q Rn RTS y U x SWV X .

A subset O of Rn is open if for every x Q O there exists an Y 0 such that B L H , x I


is contained in O. Intuitively this means that at every point in O there is a little bit
of room inside O to move around in any direction you like. An open neighbourhood
of x is any open set containing x.
A subset C of Rn is closed if its complement Rn U C is open. This definition
is equivalent to the following: C is closed if and only if for every sequence of
points x1 , x2 , . . . , xn , . . . that converges to a point x in Rn , the limit x is contained
in C. Loosely speaking, closed means closed under taking limits. An example
of a closed set is the closed ball of radius about a point x, which is defined as the
collection of all points y whose distance to x is less than or equal to ,
B H , x I/N7O y Q Rn RTS y U x SBZ X .

Closed is not the opposite of open! There exist lots of subsets of R n that are
neither open nor closed, for example the interval [ 0, 1 I in R. (On the other hand,
there are not so many subsets that are both open and closed, namely just the empty
set and Rn itself.)
A subset A of Rn is bounded if there exists some R Y 0 such that S x S Z R
for all x in A. (That is, A is contained in the ball B H R, 0 I for some value of R.) A
compact subset of Rn is one that is both closed and bounded. The importance of the
notion of compactness, as far as these notes are concerned, is that the integral of
a continuous function over a compact subset of Rn is always a well-defined, finite
number.
Exercises
A.1. Parts (iii) and (iv) of this problem require the use of an atlas (or the Web; see e.g.
). Let X be the surface of the earth, let Y be the real line and let

\^]_]a`bdc_cAegfa]ghjiAegf_kjfj]^kjf:lnmpo^iaq

128

A. SETS AND FUNCTIONS

f : X r Y be the function which assigns to each x s X its geographical latitude measured


in degrees.
(i) Find f t X u .
(ii) Find f v 1 t 0 u , f v 1 t 90 u , f v 1 txw 90 u .
(iii) Let A be the contiguous United States. Find f t A u . Round the numbers to whole
degrees.
(iv) Let B y f t A u , where A is as in part (iii). Find (a) a country other than A that
is contained in f v 1 t B u ; (b) a country that intersects f v 1 t B u but is not contained
in f v 1 t B u ; and (c) a country in the northern hemisphere that does not intersect
f v 1 t Bu .
A.2. Let S t n uzy
(i) S t 0 uy
(ii) S t n 
u y

0 { i { j { n t i | j u . Prove the following assertions.


0 and S t n | 1 uy S t n u}| 32 t n | 1 uAt n | 2 u .
1
2 n t n | 1 u~t n | 2 u . (Use induction on n.)

A.3. Prove that the open ball B :t , x u is open. (This is not a tautology! State your
reasons as precisely as you can, using the definition of openness stated in the text. You will
need the triangle inequality y w x y w z | z w x .)
A.4. Prove that the closed ball is B t , x u is closed. (Same comments as for Exercise A.3.)
A.5. Show that the two definitions of closedness given in the text are equivalent.
in

A.6. Complete the following table. Here S n v


that is the set of vectors of length 1.

Rn ,

w 3, 5
w 3, 5 u
w 3, u
txw 3, u
B t , x u
B _t , x u
Sn v 1

3
xy-plane in
R n
unit cube 0, 1

denotes the unit sphere about the origin

closed?

bounded?

compact?

yes

yes

yes

APPENDIX B

Calculus review
This appendix is a brief review of some single- and multi-variable calculus
needed in the study of manifolds. References for this material are [Edw94], [HH02]
and [MT03].
B.1. The fundamental theorem of calculus
Suppose that F is a differentiable function of a single variable x and that the
derivative f F is continuous. Let a, b be an interval contained in the domain
of F. The fundamental theorem of calculus says that
b
f t dt F b F a .
(B.1)
a

There are two useful alternative ways of writing this theorem. Replacing b with x
and differentiating with respect to x we find
x
d
f t dt f x .
(B.2)
dx a
Writing g instead of F and g instead of f and adding g a to both sides in formula
(B.1) we get
x
g x 6 g a z
(B.3)
g t dt.
a

Formulas (B.1)(B.3) are equivalent, but they emphasize different aspects of the
fundamental theorem of calculus. Formula (B.1) is a formula for a definite integral:
it tells you how to find the (signed) surface area between the graph of the function
f and the x-axis. Formula (B.2) says that the integral of a continuous function is
a differentiable function of the upper limit; and the derivative is the integrand.
Formula (B.3) is an integral formula, which expresses the function g in terms of
the value g a and the derivative g . (See Exercise B.1 for an application.)
B.2. Derivatives
Let 1 , 2 , . . . , m be functions of n variables x 1 , x2 , . . . , xn . As usual we write

x1
1 x
x2
2 x
x /
,
x . ,
..
..

xn
m x
and view x as a single map from Rn to Rm . (In calculus the word map is often
used for vector-valued functions, while the word function is generally reserved
129

130

B. CALCULUS REVIEW

for real-valued functions.) We say that is continuously differentiable if the partial


derivatives
i x he j i x
i
x lim
(B.4)

x j
h
h 0
are well-defined and continuous functions of x for all i 1, 2, . . . , n and j 1,
2, . . . , m. Here



1
0
0

0
1
0

0
0
0

e1
.. , e2
.. , . . . , en
..
.
.
.
0
0
0
0
0
1
n
are the standard basis vectors of R . The (total) derivative or Jacobi matrix of at x
is then the m n-matrix

1
x
.
.
.
x

x 1
x n
.
.
..
.. .
D x

m
m
. . . xn x
x 1 x

If v is any vector in Rn , the directional derivative of along v is defined to be the


vector D x v in Rm , obtained by multiplying the matrix D x by the vector v.
For n 1 is a vector-valued function of one variable x, often called a path
or (parametrized) curve in Rm . In this case the matrix D x consists of a single

column vector, called the velocity vector, and is usually denoted


simply by x .

For m 1 is a scalar-valued function of n variables and D x is a single

row vector. The transpose matrix of D x is therefore a column vector, usually

called the gradient of :


T
D x grad x .

The directional derivative of along v can then be written as an inner product,


D x v grad x ; v. There is an important characterization of the gradient,
is based on the identity a b a b cos . Here 0 is the angle
which
subtended by a and b. If v is a unit vector ( v 9 1), then
D x v grad x v grad x ^ cos ,

where is the angle between grad x and v. So D x v takes on its maximal

value if cos 1, i.e. 0. This means


that v points in the same direction as
grad x . Thus the direction of the vector grad x is the direction of steepest

ascent, i.e. in which increases fastest, and the magnitude


of grad x is equal

along
to the directional derivative D x v, where v is the unit vector pointing

grad x .

Frequently
a function is not defined on all of Rn , but only on a subset U. We
must be a little careful in defining the derivative of such a function. Let us assume
that U is an open set. Let : U Rm be a function defined on U and let x U.
Because U is open, there exists 0 such that the points x te j are contained
in U for t . Therefore i x te j is well-defined for t and
thus it makes sense to ask whether the partial derivatives (B.4) exist. If they do,
for all x U and all i and j, and if they are continuous, the function is called
continuously differentiable or C 1 .

B.3. THE CHAIN RULE

131

If the second partial derivatives


2 i
x
x j xk
exist and are continuous for all x U and for all i 1, 2, . . . , n and j, k 1, 2, . . . ,
m, then is called twice continuously differentiable or C 2 . Likewise, if all r-fold
partial derivatives
r i
x
x j1 x j2 :: x jr

exist and are continuous, then is r times continuously differentiable or C r . If is Cr


for all r 1, then we say that is infinitely many times differentiable, C , or smooth.
This means that can be differentiated arbitrarily many times with respect to any
of the variables.
Let us now review some of the most important facts concerning derivatives.
B.3. The chain rule
Recall that if A, B and C are sets and : A B and : B C are functions,
we can apply after to obtain the composite function x / x .


B.1. T HEOREM (chain rule). Let U Rn and V Rm be open and let : U V
and : V Rk be Cr . Then is Cr and

for all x U.

D x /

D x A D x

Here D x A D x denotes the composition or the product of the two matrices D x and D x .

B.2. E XAMPLE . In the one-variable case n m k 1 the derivatives D and


D are 1 1-matrices x A and y A , and matrix multiplication is ordinary
usual chain
the
rule
multiplication, so we get

x / x A x .

B.3. E XAMPLE . If n k 1, then is a real-valued function of one variable x, so D is a 1 1-matrix containing the single entry . Moreover,

d 1
x

dx

..
y . . . y
y  ,
D x ;
and
D y y
.
m
1

d m
dx x

so by the chain rule

d
x /
dx

D x A D x

yi x A

i 1

di
x .
dx

(B.5)

This is perhaps the most important special case of the chain rule. Sometimes we
are sloppy and abbreviate this identity to
d

dx

di
.
dx

yi

i 1

132

B. CALCULUS REVIEW

An even sloppier, but nevertheless quite common, notation is


d
dx

di
.
dx

yi

i 1

In these notes we frequently use the so-called pullback notation. Instead of


we often write , so that x stands for x A . Similarly,  yi _ x
stands for y i x A . In this notation we have
d
dx

di
.
yi dx

i 1

(B.6)

B.4. The implicit function theorem


Rm

be a continuously differentiable function defined on an open


Let : W
subset W of Rn m . Let us think of a vector in Rn m as an ordered pair of vectors
u, v with u Rn and v Rm . Consider the equation

u, v

0.

Under what circumstances is it possible to solve for v as a function of u? The


answer is given by the implicit function theorem. We form the Jacobi matrices of
with respect to the u- and v-variables separately,

1
1
1
1
.
.
.
.
.
.
u 1
u n
v 1
v m
..
..
..
..
Du
,
D

v
.
. .
.
.
m . . . m
m . . . m
u
u n
v
v m
1

Observe that the matrix Dv is square. We are in business if we have a point


u0 , v0 at which is 0 and Dv is invertible.
B.4. T HEOREM (implicit function theorem). Let : W Rm be Cr , where W
is open in Rn m . Suppose that u0 , v0 W is a point such that u0 , v0
0 and

Dv u0 , v0 is invertible. Then there are open neighbourhoods U R n of u0 and


V Rm of v0 such that for each u U there exists a unique v
f u V satis
fying u, f u A
0. The function f : U V is C r with derivative given by implicit

differentiation:
D f u

Dv u, v A

Du u, v v

f u

for all u U.
This is well-known for m n 1, when is a function of two real variables
u , v , then for u close to u and v close to
u, v . If v 0 at a certain point
0 0
0

v0 we can solve the equation u, v


0 for v as a function v
f u of u, and

u
f
.
P v
Now let us take to be of the form u, v
g v
u, where g : W R n is a

n
given function with W open in R . Solving u, v
0 here amounts to inverting

the function g. Moreover, Dv


Dg, so the implicit function theorem yields the

following result.

B.5. THE SUBSTITUTION FORMULA FOR INTEGRALS

133

B.5. T HEOREM (inverse function theorem). Let g : W


R n be continuously
n
differentiable, where W is open in R . Suppose that v0 W is a point such that Dg v0
is invertible. Then there is an open neighbourhood U R n of v0 such that g U is an
open neighbourhood of g v0 and the function g : U g U is invertible. The inverse
g 1 : V U is continuously differentiable with derivative given by
Dg

Dg v

for all v V.
Again let us spell out the one-variable case n 1. Invertibility of Dg v 0
simply means that g v0 0. This implies that near v 0 the function g is strictly
monotone increasing (if g d v0 0) or decreasing (if g d v0 0). Therefore if I is
a sufficiently small open interval around u 0 , then g I is an open interval around
g u0 and the restricted function g : I g I is invertible. The inverse function
has derivative
1
,
g 1 u /
g v
with v

u .

B.6. E XAMPLE (square roots). Let g v v 2 . Then g v0 0 whenever v0


0. For v0 0 we can take I 0, . Then g I 0, , g 1 u u, and
g 1 u 9 1 n 2 u . For v0 0 we can take I ~W , 0 . Then g I 0, ,
g 1 u u, and g 1 u 1 n 2 u . In a neighbourhood of 0 it is not
possible to invert g.
B.5. The substitution formula for integrals
Let V be an open subset of Rn and let f : V R be a function. Suppose we
want to change the variables in the integral V f y dy. (This is shorthand for an
n-fold integral over y 1 , y2 , . . . , yn .) This means we substitute y p x , where
p : U V is a map from an open U Rn to V. Under a suitable hypothesis we
can change the integral over y to an integral over x.
B.7. T HEOREM (change of variables formula). Let U and V be open subsets of R n
and let p : U V be a map. Suppose that p is bijective and that p and its inverse are
continuously differentiable. Then for any integrable function f we have


V

f y dy


U

f p xA det Dp x dx.

Again this should look familiar from one-variable calculus: if p : a, b c, d


is C 1 and has a C 1 inverse, then
 b
 d
f p x A p x dx if p is increasing,
f y dy 
 ab
a f p x A p x dx if p is decreasing.
c

 b
This can be written succinctly as cd f y dy
a f p x  p x  dx, which looks
more similar to the multidimensional case.

134

B. CALCULUS REVIEW

Exercises
h

C n 1-function,

B.1. Let g :  a, b
R be a
where n 0. Suppose a  x  b and put
x  a.
(i) By changing variables in the fundamental theorem of calculus (B.3) show that
g  x 

g  a  h 

1
0

g  a  th  dt.

(ii) Show that

 g  a  hg   a  h2 

1
0

g  a  h  t  1  g  a  th  0  h2 

g  x  

 1  t  g   a  th  dt

 1  t  g    a  th  dt.

(Integrate the formula in part (i) by parts and dont forget to use the chain rule.)
(iii) By induction on n deduce from part (ii) that
g  x  

k 0

hn 1

n!

g  k  a k
h 
k!

1
0

 1  t n g  n

 a  th  dt.

This is Taylors formula with integral remainder term.


B.2. Let x and v be constant vectors in Rn . Define c  t 
B.3. Deduce from the chain rule that D  x  v 

lim
t!

x  tv. Find c   t  .

 x  tv "  x 
.
t

B.4. According to Newtons law of gravitation, a particle of mass m 1 placed at the


origin in R3 exerts a force on a particle of mass m 2 placed at x # R3 %$ 0 & equal to
F '

Gm1 m2
( ( 3 x,
x

where G is a constant of nature. Show that F is the gradient of f  x  

Gm 1 m2 )

Rn

B.5. A function f :
*$ 0 &+
R is homogeneous of degree p if f  tx ,
x # Rn -$ 0 & and t . 0. Here p is a real constant.

( (

x .

tp f

 x  for all

(i) Show that the functions f  x, y /0 x 2  xy  )  x2  y2  , f  x, y /01 x3  y3 ,


f  x, y, z 2 x2 z6  3x4 y2 z2 4365 2 are homogeneous. What are their degrees?
(ii) Assume that f is defined at 0 and continuous everywhere. Show that p 0.
Show that f is constant if p  0.
(iii) Show that if f is homogeneous of degree p and smooth, then
n

xi xi  x 

i 1

(Differentiate the relation f  tx  

p f  x .

t p f  x  with respect to t.)

B.6. Define a function f : R


R by f  0   0 and f  x   e 3 1 7 x for x  8 0.
(i) Show that f is differentiable at 0 and that f   0  0.
(ii) Show that f is smooth and that f  n  0   0 for all n.
(iii) Plot the function f over the interval  5  x  5. Using software or a graphing
calculator is fine, but pay special attention to the behaviour near x  0.
2

B.7. Define a map from Rn 3

to Rn by

( (

2t : t

 1  en ; .

(i) Show that  t  lies on the unit sphere S n 3

about the origin.

 t  

( (
t

 19

EXERCISES

135

(ii) Show that < t = is the intersection point of the sphere and the line through the
points en and t. (Here we regard t >?< t 1 , t2 , . . . , tn @ 1 = as a point in Rn by identifying it with < t 1 , t2 , . . . , tn @ 1 , 0 = .)
(iii) Compute D < t = .
(iv) Let X be the sphere punctured at the north pole, X > S n @ 1 ACB en D . Stereographic projection from the north pole is the map : X E Rn @ 1 given by < x =F>
< xn A 1 = @ 1 < x1 , x2 , . . . , xn @ 1 = . Show that is a two-sided inverse of .
(v) Draw diagrams illustrating the maps and for n > 2 and n > 3.
(vi) Now let y be any point on the sphere and let P the hyperplane which passes
through the origin and is perpendicular to y. The stereographic projection from y
of any point x in the sphere distinct from y is defined as the unique intersection point of the line joining y to x and the hyperplane P. This defines a map
: Sn @ 1 ACB y D E P. The point y is called the centre of the projection. Write a
formula for the stereographic projection from the south pole A en and for its
inverse : Rn @ 1 E Sn @ 1 .
B.8. A map : Rn E
is even and C 1 .

Rm is called even if < A x =G>

< x = for all x in Rn . Find D < 0 = if

B.9. Let a0 , a1 , a2 , . . . , an be vectors in Rn . A linear combination niH 0 ci ai is convex if


> 1. The simplex I spanned by the ai s is the collection of all their convex linear
combinations,
niH 0 ci

IC>KJ

iH 0

iH 0

ci ai LL ci >
L

1M .

L
The standard simplex in Rn is the simplex spanned
by the vectors 0, e1 , e2 , . . . , en .
(i) For n > 1, 2, 3 draw pictures of the standard n-simplex as well as a nonstandard
n-simplex.
(ii) The volume of a region R in Rn is defined as N

dx1 dx2 OPOPO dxn . Show that

1
det A ,
Q
n! Q
where A is the n R n-matrix with columns a1 A a0 , a2 A a0 , . . . , an A a0 . (First
compute the volume of the standard simplex by repeated integration. Then map
I to the standard simplex by an appropriate substitution and apply the substitution formula for integrals.)
vol IC>

The following two calculus problems are not review problems, but the results are
needed in Chapter 9.
B.10. For x S

0 define

< x =>
T followingT assertions.
and prove the

NVU e @ t t x @
0

(i) T < x W 1 = > x < x = for all x S 0.


= >X< n A 1 = ! for TZpositive
(ii) < n
integers n.
Y
1 aW 1
@
u2 a
u du >
(iii) NVU e
.
2
2 T [
0
T

dt

T defined in Exercise B.10) by establishB.11. Calculate < n W 12 = (where is the function


ing the following identities. For brevity write >
< 12 = .
2
N U e @ s ds.
@
(ii) 2 > N U U N U e @
@
@
U
U

(i) >

x2 @ y2

dx dy.

136

B. CALCULUS REVIEW

(iii) 2 \^]

0
(iv) \^b .

(v) ced n f

]`_ re a

1 \
2g

r2

dr d.

1 h 3 h 5 hPhPhji 2n k 1 l b
for n m
2n

1.

Bibliography
[Bac06]

[BS91]
[BT82]

[Bre91]

[Bre05]

[Car94]

[Dar94]

[Edw94]

[Fla89]

[Gra98]

[GP74]

[HH02]
[MT03]
[Opr03]

[Sin01]
[Spi65]

D. Bachman, A geometric approach to differential forms, Birkhuser, Boston, MA, 2006.


A recent text, a version of which is available through the authors website
nPojoqprsjstpPuPvxw4vqyz{p}|~ojuPqzqyqsPtyqvxwn4v4s .
P. Bamberg and S. Sternberg, A course in mathematics for students of physics, Cambridge University Press, Cambridge, 1991.
R. Bott and L. Tu, Differential forms in algebraic topology, Springer-Verlag, New York, 1982.
Masterly exposition at beginning graduate level of the uses of differential forms in topology
and de Rham cohomology.
D. Bressoud, Second year calculus, Undergraduate Texts in Mathematics, Springer-Verlag, New
York, 1991.
Multivariable calculus from the point of view of Newtons law and special relativity with
coverage of differential forms.
O. Bretscher, Linear algebra with applications, third ed., Pearson Prentice Hall, Upper Saddle
River, New Jersey, 2005.
Good reference for the basic linear algebra required in these notes.
M. do Carmo, Differential forms and applications, Springer-Verlag, Berlin, 1994, translated from
the 1971 Portuguese original.
Slightly more advanced than these notes, with some coverage of Riemannian geometry, including the Gau-Bonnet theorem.
R. Darling, Differential forms and connections, Cambridge University Press, Cambridge, 1994.
Advanced undergraduate text on manifolds and differential forms, including expositions of
Maxwell theory and its modern generalization, gauge theory.
H. Edwards, Advanced calculus: A differential forms approach, Birkhuser Boston, Boston, Massachusetts, 1994, corrected reprint of the 1969 original.
One of the earliest undergraduate textbooks covering differential forms. Still recommended
as an alternative or supplementary source.
H. Flanders, Differential forms with applications to the physical sciences, second ed., Dover Publications, New York, 1989.
Written for 1960s engineering graduate students. Concise but lucid (and cheap!).
A. Gray, Modern differential geometry of curves and surfaces with Mathematica, second ed., CRC
Press, Boca Raton, FL, 1998.
Leisurely and thorough exposition at intermediate undergraduate level with plenty of computer graphics.
V. Guillemin and A. Pollack, Differential topology, Prentice-Hall, Englewood Cliffs, N.J., 1974.
Written for beginning graduate students, but very intuitive and with lots of interesting applications to topology.
J. Hubbard and B. Hubbard, Vector calculus, linear algebra, and differential forms: A unified approach, second ed., Prentice Hall, Upper Saddle River, New Jersey, 2002.
J. Marsden and A. Tromba, Vector calculus, fifth ed., W. H. Freeman, New York, 2003.
Standard multivariable calculus reference.
J. Oprea, Differential geometry and its applications, second ed., Prentice Hall, 2003.
Curves, surfaces, geodesics and calculus of variations with plenty of MAPLE programming.
Accessible to intermediate undergraduates.
S. Singer, Symmetry in mechanics. A gentle, modern introduction, Birkhuser, Boston, MA, 2001.
M. Spivak, Calculus on manifolds. A modern approach to classical theorems of advanced calculus, W.
A. Benjamin, New York-Amsterdam, 1965.
Efficient and rigorous treatment of many of the topics in these notes.
137

138

[Spi99]

BIBLIOGRAPHY

, A comprehensive introduction to differential geometry, third ed., Publish or Perish, Houston, TX, 1999.
Differential geometry textbook at advanced undergraduate level in five massive but fun to
read volumes.
[Wei97] S. Weintraub, Differential forms. A complement to vector calculus, Academic Press, San Diego,
CA, 1997.
Written as a companion to multivariable calculus texts. Contains careful and intuitive explanations of several of the ideas covered in these notes, as well as a number of straightforward
exercises.

The Greek alphabet


upper case lower case name
A
B

E
Z
H

I
K

M
N

,

o
, $

139

alpha
beta
gamma
delta
epsilon
zeta
eta
theta
iota
kappa
lambda
mu
nu
xi
omicron
pi
rho
sigma
tau
upsilon
phi
chi
psi
omega

Notation Index

dx, infinitesimal hypersurface, 26


dx, infinitesimal displacement, 25
dx I , short for dx i1 dx i2 ~ dx ik , 17
dx i , covector (infinitesimal increment), 17,
83
dx i , omit dx i , 18

, Hodge star operator, 24


29
relativistic,
k , unit cube in Rk , 58
0,
1

, orientation defined by a basis , 94


, Euclidean inner product (dot product), 8

, composition of maps, 37, 126, 131


, integral of a form
over a chain, 57
over a manifold, 106

of a form over a chain, 47


V, integral

, Euclidean norm (length), 8


, tensor multiplication, 89

, partial derivative, 130


xi
, orthogonal complement, 74

, exterior multiplication, 17, 85

ei , i-th standard basis vector of Rn , 130


f A, restriction of f to A, 126
g f , composition of f and g, 37, 126, 131

, Gamma function, 110, 135


grad, gradient of a function, 26
graph, graph of a function, 68
H n , upper halfspace in Rn , 103

A T , transpose of a matrix A, 35
A k V, set of alternating k-multilinear functions
on V, 85
A , permutation matrix, 43
A I, J , I, J-submatrix of A, 42
, Hodge star of , 24
relativistic, 29
M , integral of over a manifold M, 106
c , integral of over a chain c, 47, 57
Alt , alternating form associated to , 90

I, multi-index i 1 , i 2 , . . . , i k (usually increasing), 17


int M, interior of a manifold with boundary,
103
ker A, kernel (nullspace) of a matrix A, 73
l , length of a permutation , 34

M , volume form of a manifold M, 96

k , binomial coefficient, 19, 23


n, unit normal vector field, 95
nullity A, dimension of the kernel of A, 73

B , x , closed ball in Rn , 127


n
B
4 , x , open ball in R , 127

, orientation defined by a basis , 94

O n , orthogonal group, 76
k
M , vector space of k-forms on M, 19, 82

C r , r times continuously differentiable, 131


curl, curl of a vector field, 27

, pullback
of a form, 37, 88
of a function, 37, 132

D, Jacobi matrix of , 130


, boundary
of a chain, 59
of a manifold, 103
d, exterior derivative, 20

, Laplacian of a function, 28
I, J , Kronecker delta, 86
i, j , Kronecker delta, 51
det A, determinant of a matrix A, 31
div, divergence of a vector field, 26

Rn , Euclidean n-space, 1
rank A, dimension of the column space of A,
73
S n , unit sphere about the origin in Rn
S n , permutation group, 33
sign , sign of a permutation , 34
141

, 8, 64

142

NOTATION INDEX

SL n , special linear group, 78


Tx M, tangent space to M at x, 8, 69, 73, 74
V , dual of a vector space V, 83
V k , k-fold Cartesian product of a vector space
V, 85
voln , n-dimensional Euclidean volume, 91

x , Euclidean norm (length) of a vector x, 8


x y, Euclidean inner product (dot product) of
vectors x and y, 8
x T , transpose of a vector x, 1

Index
Page numbers in boldface refer to definitions or theorems; italic
page numbers refer to examples or applications. In a few cases
italic boldface is in order.
circle, 9, 47, 48, 5153, 55, 62, 64, 72, 115, 118,
123
closed
ball, 8, 105, 106, 110, 115, 122, 127
chain, 62, 64, 122
curve, 1, 50, 53, 62
form, 22, 2729, 40, 51, 54, 55, 117123
set, 4, 36, 47, 84, 103, 106, 107, 127, 128
codimension, 69, 73, 74, 105
cohomology, 137
column
operation, 32, 42
vector, 1, 31, 45, 69, 83, 89, 92, 130
compact, 57, 106110, 117, 120, 127
complementary, 24
configuration space, 9, 14
connected component, 12, 119
conservative, 49, 108, 109
constant form, 19, 28, 83
continuously differentiable, 130
contractible, 115, 119122
contraction, 115, 120, 121122
contravariance, 38
convex, 122
linear combination, 135
coordinate map, see chart
covariant vector, see covector
covector, 83, 84, 85, 89, 90
field, 83
Coxeter, Harold Scott MacDonald (19072003),
43
Coxeter relations, 43
critical point, 79
cross-cap, 7
cube in an open set, 59, 6062, 6465
curl, 27, 29, 65, 109
curvature, 17
curve, 1, 69
cycle, 62, 64
cylinder
formula, 116

affine space, 7, 8, 73
alternating
algebra, 82
multilinear function, 32, 85, 8790
property, 18, 19, 21
Ampre, Andr Marie (17751836), 29
Ampres Law, 29
angle
form, 23, 36, 47, 51, 64, 118, 120, 123
function along a curve, 5253
anticommutativity, 18
antisymmetric
matrix, 78
multilinear function, see alternating multilinear function
arc length, 89, 96, 101, 107
Archimedes of Syracuse (287212 BC), 3, 109
Archimedes Law, 109
atlas, 67, 69, 82
average of a function, 107
ball, see closed ball, open ball
barycentre, 107
bilinear, 85, 89
block, 31, 42, 91, 93, 96
rectangular, see rectangular block
Bonnet, Pierre (18191892), 137
boundary
of a chain, 60, 6265
of a manifold, 27, 103, 104111, 118
bounded, 47, 106, 127
Brouwer, Luitzen Egbertus Jan (18811966), 113,
122
Brouwers fixed point theorem, 113, 122
Cartan, Elie (18691951), 17
Cartesian product, 10, 85, 116, 125
Cartesius, Renatus, see Descartes, Ren
centroid, 107
chain, 59, 6065
chart, 67, 69, 72
143

144

with base M, 116


dAlembert, Jean Le Rond (17171783), 29
dAlembertian, 29
de Rham, Georges (19031990), 137
degenerate chain, 61, 64
degree
of a form, 17
of a homogeneous function, 134
of a multi-index, 17
degrees of freedom, 9, 14
Descartes, Ren (15961650), 10, 85, 125
determinant, 31, 3235, 4042, 43, 78, 85, 86,
9194
differential
equation, 15
form, see form
dimension, 111, 69
Dirichlet, Lejeune (18051859), 110
Dirichlet integral, 110
disconnected, 12, 108
divergence, 26, 29, 65, 108, 109
domain, 106, 108, 109
of a function, 125
dot product, see inner product
dual
basis, 84, 86, 87, 89
vector, see covector
space, 83
electromagnetic wave, 29
electromagnetism, 29
element
of arc length, 89, 96, 101, 123
of surface area, 96
embedding, 67, 68, 70, 71, 7778, 81, 88, 9698,
100, 101, 103, 104, 106, 107
Euclid of Alexandria (ca. 325265 BC), 1, 67,
69, 91, 110, 114, 125, 127
Euclidean
motion, 91
plane, 125
space, 1, 67, 69, 110, 114, 127
volume, 91

, 109
even
map, 135
permutation, 34, 43
exact form, 22, 28, 4952, 64, 117120, 123
exterior
algebra, 82
derivative, 20, 21, 25, 27, 60, 63
on a manifold, 82
differential calculus, 17
product, see product of forms
Faraday, Michael (17911867), 29
Faradays Law, 29
fibre, see level set

INDEX

fixed point, 113


flux, 26, 99, 108, 109
form
as a vector-eating animal, 88
closed, see closed form
exact, see exact form
on a manifold, 82, 88
on Euclidean space, 17, 1830, 88
volume, see volume form
free space, 29
Freedman, Michael (1951), 121
function, 125, 130
functional, see covector
fundamental theorem of calculus, 27, 47, 49,
52, 64, 129, 134
in Rn , 49, 63, 108
Gamma function, 110111, 135
Gau, Carl Friedrich (17771855), v, 29, 63, 65,
95, 108, 137
Gau map, 95
Gau Law, 29
graded commutativity, 18, 19
gradient, 26, 28, 49, 65, 7479, 96, 100, 101, 108,
110, 130
Gram, Jorgen (18501916), 92
Gram-Schmidt process, 92
graph, 68, 70, 73, 97, 100, 103, 129
Gramann, Hermann (18091877), 82
gravitation, 54, 134
Greek alphabet, 139
Green, George (17931841), v, 63, 65, 110
Greens theorem, 65, 110
halfspace, 103
Hodge, William (19031975), 23, 25, 28, 29, 38
Hodge star operator, 23, 25, 28, 38, see also relativity
homogeneous function, 28, 78, 122, 134
homotopy, 114
formula, 117
of curves, 115
of loops, 115, 119, 121
hypersurface, 26, 69, 95101, 106, 108, 120
increasing multi-index, 19, 24, 28, 41, 86, 87,
see also complementary
index of a vector field, 55
inner product, 8, 84, 85, 98, 130
of forms, 28
integrability condition, 22
integral
of a 1-form over a curve, 47, 48, 49, 51
of a form
over a chain, 57, 5859, 6365, 106
over a manifold, 106, 107, 108, 117, 119,
120, 122
inversion, 34
inward pointing, 106

INDEX

k-chain, see chain


k-cube, see cube
k-form, see form
k-multilinear function, see multilinear function
Klein, Felix (18491925), 5
Klein bottle, 5
Kronecker, Leopold (18231891), 51
Kronecker delta, 51
Laplace, Pierre-Simon (17491827), 28
Laplacian, 28
Leibniz, Gottfried Wilhelm von (16461716), 20,
21, 40
Leibniz rule
for forms, 21, 40
for functions, 20
length
of a permutation, 34, 4243
of a vector, 8, 76, 79, 95, 106, 114, 128
level
curve, 74
hypersurface, 74
set, 73, 126
surface, 74
lightlike, 29
line segment, 91
local representative, 82, 88, 89, 96
Lotka, Alfred (18801949), 15, 78
Lotka-Volterra model, 15, 78
manifold, 115, 69, 7079
abstract, 913, 69
given explicitly, 9, 72
given implicitly, 7, 72
with boundary, 27, 103, 104111, 118
map, 125, 130
Maxwell, James Clerk (18311879), 29
mean of a function, 107
measurable set, 36
Milnor, John (1931), 121
minimum, 75
Minkowski, Hermann (18641909), 29
Minkowski
inner product, 29
space, 29
Mbius, August (17901868), 4, 105
Mbius band, 4, 105
multi-index, 17, 126, see also increasing multiindex
multilinear
algebra, 17
function, 85, 89
n-manifold, see manifold
naturality of pullbacks, 38, 48, 58
Newton, Isaac (16431727), 54, 134, 137
norm of a vector, 8
normal vector field, see unit normal vector field

145

odd permutation, 34, 43


open
ball, 127
neighbourhood, 127
set, 127, 128
is a manifold, 70
orientation
of a boundary, 106, 108
of a hypersurface, 95, 100
of a manifold, 88, 95, 108
of a vector space, 91, 94, 100
preserving, 48, 57, 9597, 101, 106, 107
reversing, 48, 57, 95
orthogonal
complement, 74, 95
group, 76, 78
matrix, 76, 91, 92
operator, 28
projection, 93
orthonormal, 28, 91
outward pointing, 106
outward-pointing, 110
pair of pants, 105
paraboloid, 4
parallelepiped, see block
parallelogram, 13, 14, 91, 100
parametrization, 9, 57, 67
parametrized curve, 13, 47, 49, 5355, 7778,
130
partial differential
equation, 22
operator, 20
path, see parametrized curve
pentagon, 14
Perelman, Grigori (1966), 121
periodic function, 123
permutation, 33, 34, 35, 41, 4243, 85, 100
group, 33, 43
matrix, 43
pinch point, 7
plane curve, 1, 13, 69
Poincar, Jules Henri (18541912), 119, 121
Poincar
conjecture, 121
lemma, 119
potential, 49, 53, 54, 108
predator, 15
prey, 15
product
of forms, 19, 27
on a manifold, 82
of permutations, 34, 43
of sets, see Cartesian product
product rule, see Leibniz rule
projective plane, 5
pullback
of a form

146

on a manifold, 88, 97, 106, 114, 116, 117


on Euclidean space, 37, 4042, 47, 48, 57,
58, 82, 88
of a function, 132
punctured
Euclidean space, 78, 114, 119121
plane, 47, 51, 64, 77, 115, 119, 120
quadrilateral, 11
rectangular block, 18, 57, 65
regular value, 73, 7479, 96, 100, 105
relativity, 10, 29, 137
reparametrization
of a curve, 47, 48, 50, 58
of a rectangular block, 57, 58
restriction of a map, 47, 106, 107, 110, 126
retraction, 113
Riemann, Bernhard (18261866), 137
rigid body, 10
row
operation, 42
vector, 1, 33, 45, 74, 83, 89, 130
saddle point, 75
Schmidt, Erhard (18761959), 92
sign of a permutation, 34, 4243
simple permutation, 43
simplex, 135
simply connected, 119, 121
singular
cube, see cube
value, 73, 7476, 78
singularity, 2, 8, 13, 14, 67, 71, 105
Smale, Stephen (1930), 121
smooth
curve, 69
function or map, 131, 134
hypersurface, 69
manifold, 4
point, 2
surface, 69
solution curve, 9, 15
space curve, 1
space-time, 29
spacelike, 29
special linear group, 78
sphere, 4, 8, 10, 11, 14, 64, 67, 75, 79, 96, 100,
105, 106, 110, 113, 114, 119121, 128, 134
spherical
coordinates, 44
pendulum, 9
standard
basis of Rn , 130
orientation, 94
simplex, 135
star-shaped, 122
state space, see configuration space
steepest ascent, 75, 130

INDEX

stereographic projection, 72, 135


Stirling, James (16921770), 111
Stirlings formula, 111
Stokes, George (18191903), v, 47, 63, 65, 107,
109
Stokes theorem
classical version, 65, 109
for chains, 47, 63, 64
for manifolds, 103, 108, 116, 118, 120
submanifold, 69
surface, 4, 14, 69, 81, 109
area, 9, 17, 96, 107, 109, 129
symmetric
bilinear function, 85
matrix, 76, 79
tangent
hyperplane, 69
line, 1, 13, 69, 70, 77
plane, 14, 69, 78
space, 1, 8, 14, 69, 70, 7376, 78, 88, 95, 96,
106
vector, 48, 69, 79, 88, 106
Taylor, Brook (16851731), 134
Taylors formula, 134
tensor product, 17, 85, 89
timelike, 29
topological manifold, 4
torus, 4, 68, 106
trace, 78
trajectory, 15, 78
transformation law, 82
transpose of a matrix or a vector, 1, 26, 35, 74,
83, 130
transposition, 42
unit
circle, see circle
cube, 35, 58, 65, 128
interval, 35, 58, 91, 114, 116
normal vector field, 95, 98, 99, 101, 106, 108
110, 120
sphere, see sphere
square, 35, 60, 62
vector, 51, 130
vector, see also column, length, row, unit, tangent
field, 24, 28, 29, 48, 55, 65, 98, 108, 109, see
also conservative, curl, divergence, gradient, index, potential, unit normal
Volterra, Vito (18601940), 15, 78
volume
change, 36, 42
element, 17, 96
Euclidean, see Euclidean volume
form, 65, 96, 97, 103
of a hypersurface, 99
on Rn , 18, 24, 42

INDEX

on a hypersurface, 100, 110, 120


of a block, 18, 31, 91, 92, 93
of a manifold, 89, 107, 109
of a simplex, 135
wave operator, 29
wedge product, 17, 85, 89
winding number
of closed curve, 53, 5455, 118, 120
of hypersurface, 120
work, 17, 25, 48, 49, 61, 108, 109
zero of a vector field, 25

147

You might also like